This study aims to investigate and compare the adsorption behaviour of pine needle biochar (PNB) and H2O2-oxidized PNB (OPNB) in eliminating Cu(II) from acid mine drainage. The PNB and OPNB adsorbents undergo comprehensive characterization through various techniques (BET, FTIR, SEM, and pHPZC). A central composite design was employed for designing experiments and optimizing the impact of process factors (metal concentration, adsorbent doses, contact time, and pH) on adsorption capacity. Pseudo-first-order, pseudo-second-order, and intra-particle diffusion kinetics models as well as Langmuir, Freundlich, and Temkin isotherm models were used to analyze the experimental data. Langmuir isotherm best fit (R2 > 0.99) the experimental data and adsorption capacities of 29.49 and 102.04 mg/g, were determined for PNB and OPNB, respectively. Under optimized experimental conditions, desorption studies revealed the reusability of OPNB about 80% even after four cycles. Fixed-bed column experiments were conducted at ambient temperature with an initial Cu(II) concentration of 125 mg/L and 5.0 g of adsorbent, utilizing a flow rate of 1 mL/min for both PNB and OPNB. These results indicate that oxidized biochar, synthesized for Cu(II) remediation, not only addresses Himalayan pine needle concerns sustainably but also exhibits potential applicability for removing other metal ions from aqueous environments.

  • Resource in the form of biochar can be generated from Himalayan Forest waste pine needles.

  • The application of H2O2 oxidation resulted in an increment of oxygen content in the Himalayan pine needle biochar.

  • The adsorption isotherm of the H2O2-modified pine needle biochar exhibited the best fit with the Langmuir model.

  • H2O2-oxidized pine needle biochar demonstrates efficacy as an effective adsorbent for the removal of copper.

Acid mine drainage (AMD) causes environmental problems due to low stream pH and high concentrations of heavy metals (HMs) (Shane et al. 2021). When these HMs react with dissolved oxygen, they form harmful metal oxides, endangering aquatic life and the food chain (Kefeni et al. 2017). Copper (Cu) concentration in the AMD ranged from 3 to 138 mg/L which significantly contributes to environmental pollution (Edraki et al. 2005; Ayora et al. 2016; Rodríguez-Galán et al. 2019). The chemistry for the release of Cu from AMD is explained subsequently. As per Equations (1)–(4), natural acid drainage arises within mine materials like waste rocks and tailings. This occurs when residual sulphide minerals, like pyrite (FeS2), oxidize in the presence of air, water, and specific bacteria (Larsson et al. 2018; Shane et al. 2021). This process releases protons (H+) (Equation (1)), subsequently lowering pH levels. Notably, iron and sulphur-oxidizing bacteria play a crucial role in catalyzing these reactions (Equations (2)–(4)) at low pH, significantly accelerating the reaction rates.
(1)
(2)
(3)
(4)
As a result of above reactions is produced by the overall reaction as follows (Equation (5)) (Shane et al. 2021):
(5)
Subsequently, this acid () becomes capable of dissolving additional minerals, releasing the metals they contain. Notably, one of the primary metals liberated through this chemical process is copper (Equation (6)) (Shane et al. 2021).
(6)

Elevated copper levels harm both plant and human health. In plants, it hinders growth, disrupts cellular function, and affects photosynthesis (Simate & Ndlovu 2014; Masindi et al. 2018). In humans, it leads to conditions like anaemia, liver and kidney damage, and stomach/intestinal irritation. Additionally, copper, along with HMs like cadmium, lead, and zinc, poses significant threats to aquatic life (Jiwan & Ajay 2011; Rodríguez-Galán et al. 2019). Acute exposure to these metals, with high short-term concentrations, can result in direct mortality, while chronic exposure, characterized by prolonged low concentrations, can lead to stunted growth, reduced reproduction, and deformities (Jiwan & Ajay 2011; Simate & Ndlovu 2014; Shane et al. 2021).

To date, various traditional methods, including chemical precipitation, ion exchange (Zhang et al. 2013), coagulation, membrane filtration, zero valent iron and adsorption have been employed to eliminate HMs from water (Al-Saydeh et al. 2017; Crane & Sapsford 2018). Among these approaches, adsorption distinguishes itself as a favoured option owing to its operational simplicity and notable treatment effectiveness. A range of adsorbents, including activated carbon, ion exchangers, cost-effective alternatives (such as biomass based sorbents), and biochar have been developed by various researchers for the effective HMs removal from contaminated water (Anastopoulos et al. 2019; Qiu et al. 2021).

Biochar, a carbon-rich organic material produced through biomass pyrolysis in low-oxygen conditions (EBC 2022), is an affordable and effective option for removing HM ions from polluted water (Qiu et al. 2021). Its negative surface charge, high specific surface area (SSA), porosity, and charge density make it compatible and selective for the adsorption of HMs (Deng et al. 2017; Qiu et al. 2021). Biochar application in water enhances the immobilization of HMs through a combination of electrostatic and non-electrostatic forces (Deng et al. 2017). A number of studies have been carried out by the various researchers on the removal of Cu(II) by biochar in aqueous solutions (Chen et al. 2011; Choudhary et al. 2020). Biochar derived from corn straw exhibited maximum Cu(II) adsorption capacity of 12.5 mg/g (Chen et al. 2011). The maximum sorption capacity for Cu(II) removal using pine needle biochar (PNB) was reported as 40.4 mg/g (Choudhary et al. 2020). On the other hand, researchers have reported exceptionally high adsorption capacity upto 373 mg/g for Cu(II) adsorption using activated carbon prepared from walnut shell (Shu et al. 2017).

In comparison to commercial activated carbon, biochar exhibits a relatively low adsorption capacity for HMs. Consequently, various modification methods were employed to enhance its HMs removal capabilities. Common methods involve the modification through H3PO4 (Peng et al. 2017), KOH (Trakal et al. 2014; Bashir et al. 2018), FeCl3 for chemical modification, and physical activation techniques such as steam (Kwak et al. 2019) and carbon dioxide activation (Chang et al. 2000). However, these activation processes often require high temperatures, particularly for steam, and carbon dioxide activation (typically > 800 °C), leading to increased adsorbent costs (San Miguel et al. 2003).

In contrast, H2O2 is a relatively inexpensive and clean alternative, unlikely to cause secondary pollution as its decomposition products are H2O and O2 (Wang & Liu 2018). On the other hand, H2O2 oxidation enhances biochar physicochemical properties by introducing oxygen-containing functional groups (–COOH and –OH), expanding its SSA and pore size (Fan et al. 2018). Key factors in the adsorption process include π–π electron donor-receptor interactions, hydrogen bonding effects, hydrophobic distribution, and pore filling effects (Bhattacharyya 2016; Wang & Liu 2018). Additionally, HM adsorption on biochar primarily occurs through cation-π interactions, surface complexation, and ion exchange. The formations of stable complexes or competition for sites with multifunctional groups are crucial mechanisms for inorganic pollutant interactions with HMs on biochar (Fan et al. 2018). These relationships are vital for analyzing the impact of H2O2 concentration on metal removal.

Accordingly, in this study, Himalayan PNB was oxidized with H2O2 to enhance its performance and specificity in Cu(II) uptake. This study used the response surface methodology (RSM), employed through Design-Expert software (version 11.0), for designing experiments and statistical optimization. Additionally, the research considered four parameters (Cu(II) concentration, adsorbent dose, contact time and pH), various adsorption isotherms, and kinetic models for analyzing the equilibrium data.

Biochar preparation and modification

PNB was prepared from the Himalayan pine needles (PNs) according to the method adopted elsewhere (Bashir et al. 2022). The oxidized biochar was prepared at ambient temperature (25 ± 2 °C), by mixing 3 g of PNB sample with 20 mL of 3, 10, 20, and 30% H2O2 solutions (Wang & Liu 2018). The resulting oxidized biochar was washed with deionized water (DI) until a neutral pH was achieved, followed by drying at 105 °C for 24 h to remove moisture. The H2O2-oxidized biochar was labelled as OPNB_3, OPNB_10, OPNB_20, and OPNB_30 after treatment with 3, 10, 20, and 30% H2O2 solution, respectively. Biochar and the oxidized biochar samples were kept in a sealed container for further experiments.

Synthetic wastewater

Copper (Cu(II)) stock solution of 1,000 mg/L was prepared by dissolving 3.804 g of Cu(NO3)2·3H2O in 1 L deionized (DI) water. Solutions with varying concentrations of Cu(II) determined through central composite design (CCD) as outlined in Table 1 were subsequently prepared using the stock solution.

Table 1

Parameters of the batch experiment

ParametersUnitsPNB
OPNB
− 1+ 10+ α− 1+ 10+ α
Copper Concentration (A) (mg/L) 50 125 87.5 12.5 162.5 50 125 87.5 12.5 162.5 
Adsorbent Dose (B) (g/L) 0.4 0.8 0.6 0.2 0.4 0.8 0.6 0.2 
Time (C) (min) 360 960 660 60 1,260 360 960 660 60 1,260 
pH (D)  4.5 1.5 7.5 4.5 1.5 7.5 
ParametersUnitsPNB
OPNB
− 1+ 10+ α− 1+ 10+ α
Copper Concentration (A) (mg/L) 50 125 87.5 12.5 162.5 50 125 87.5 12.5 162.5 
Adsorbent Dose (B) (g/L) 0.4 0.8 0.6 0.2 0.4 0.8 0.6 0.2 
Time (C) (min) 360 960 660 60 1,260 360 960 660 60 1,260 
pH (D)  4.5 1.5 7.5 4.5 1.5 7.5 

Batch experiments

During the study, Cu(II) batch adsorption experiments were conducted with synthetic wastewater (SWW) that mimicked real AMD. Table 1 provides a summary of the process parameters used in the batch tests. The copper concentrations were varied within the range of 12.5–162.5 mg/L (12.5, 50, 87.5, 125, and 162.5 mg/L), which is consistent with Cu(II) concentrations reported in AMD in various studies (ranging from 3 to 138 mg/L) (Edraki et al. 2005; Ayora et al. 2016; Rodríguez-Galán et al. 2019). The pH values were adjusted to 1.5, 3, 4.5, 6.0, and 7.5 to cover a broad spectrum of experimental conditions. Additionally, different contact times ranging from 60 to 1,260 min were examined (Table 1). The adsorbent dosages used in the study ranged from 0.2 to 1 g/L for both the adsorbents (PNB and OPNB).

The adsorption experiments were conducted in triplicate using 50-mL Erlenmeyer flasks, each containing 20 mL of Cu(II) ion solutions with varying initial concentrations and employing PNB and OPNB adsorbents. pH of the solutions at initial stage of experiments was maintained using either 0.01 M HCl or 0.01 M NaOH. To facilitate effective contact between the adsorbent and adsorbate, the samples were mixed at 200 (revolutions per minute) RPM using an orbital shaker (Tarson, 3050, Spinix, MC01) at an ambient temperature of 25 ± 2 °C. All the experiments were conducted in triplicate to minimize the errors. Subsequently, the supernatant was subjected to filtration using a 0.2-μm syringe filter and subsequently diluted with 2% HNO3 for analysis.

Design of experiments

The batch adsorption experiments employed the CCD method, which included 2k factorial runs, 2k axial runs, and a midpoint (n0). Independent variables were coded as −α, −1, 0, +1, and +α. Four parameters namely copper concentration, adsorbent doses (PNB/OPNB), contact time and pH were considered for the optimization. The experiment consists of 16 factorial, 8 axial, and 6 centre duplicates, accounted for 30 experiments according to the following formula:
(7)

Here, N indicates the total number of experiments, k indicates the number of factors or parameters, n0 indicates the number of central point replicates.

RSM was employed to assess how various factors collectively influenced the adsorption capacity (Q) of PNB and OPNB. These parameters were determined using the following equations:
(8)

Here, Ci indicates the initial concentration of Cu(II) (mg/L), Ce indicates the final concentration of Cu(II) (mg/L), V indicates the volume of the aqueous solution (L), M indicates the absorbent mass (g).

Adsorption isotherm models

To understand the Cu(II) adsorption mechanism using PNB and OPNB as an adsorbent, three models, namely, Langmuir, Freundlich, and Temkin adsorption isotherms were investigated (Bashir et al. 2022). Isotherm models assist in designing optimal adsorption conditions for various applications.

Langmuir isotherm

Langmuir isotherm (Equation (9)), which describes monolayer adsorption explains reversible adsorption of species on surfaces, treating the adsorbate like an ideal gas at isothermal conditions. It quantifies adsorbate amounts in relation to concentration at a given temperature (Langmuir 1918). KL (L/mg), known as the Langmuir adsorption constant and describing the affinity between the adsorbate and adsorbent, provided insights into the strength of the adsorption process. Plotting of Ce/qe versus Ce was used to find out the slope and intercept from the linearized Equation (10) that yielded the values of qm and KL.
(9)
(10)
where qm indicates the maximum monolayer adsorption capacity (mg/g).

Freundlich isotherm

The Freundlich isotherm, which describes heterogeneity of the surface and multilayer adsorption phenomena given by entirely empirical Equation (11) (Karthikeyan et al. 2005). The constant KF ((mg/g) (L/mg)1/n) in Equation (11) serves as a crucial factor in characterizing the interaction between the adsorbate and adsorbent. The KF value contributes to understanding the adsorption capacity of the material, highlighting the quantity of adsorbate that can be adsorbed per unit mass of the adsorbent, as well as the strength of the interaction. It can be expressed using Equation (11), and upon simplification, its linearized form is represented by Equation (12).
(11)
(12)

Here, n indicates the dimensionless constant (signifying the adsorption intensity).

Temkin isotherm

The Temkin isotherm model, as expressed by Equation (13), stands out for its unique characteristic of ensuring a uniform distribution of binding energies across the adsorbent surface (Gebrewold et al. 2019). The determination of the Temkin constants (KT, BT) was accomplished by plotting qe vs ln(Ce), and calculating these constants from the slope and intercept of the resulting graph (Gebrewold et al. 2019). The Temkin coefficient (KT) plays a pivotal role in characterizing the isotherm and elucidating the behaviour of the adsorption system. It reflects the strength and nature of interactions between the adsorbate and adsorbent, providing valuable information about the surface heterogeneity and the binding energy of the molecules.
(13)
where BT indicates the Temkin adsorption potential (kJ/mol).

Kinetic study

In general adsorption kinetics is a graphical representation, which depicts the rate at which a solute is either retained or released from an aqueous solution to the solid-phase interface, considering factors such as adsorbate concentration, adsorbent dosage, temperature, and pH (Fakhri 2017). The adsorption process involves two primary mechanisms: physical adsorption, or physisorption, which arises from weak forces like van der Waals interactions, and chemical adsorption, or chemisorption, where a strong bond forms between the adsorbate and the adsorbent, involving electron transfer.

The kinetic study of the data obtained from copper sorption batch test was analyzed by using the following kinetic models:

Pseudo-first-order kinetics model

The pseudo-first-order (PFO) model, also referred to as the Lagergren model, characterizes the adsorption of an adsorbate onto an adsorbent by employing a first-order reaction (Equation (14)) mechanism (Gebrewold et al. 2019). The linear form of the PFO is presented in Equation (15)). According to the PFO model, the adsorption process involves mechanisms, such as chemical and electrostatic interactions occurring among functional groups on the adsorbents surface and the adsorbate molecules. The rate constant of (PFO) adsorption, denoted as k1(L/min), is a crucial parameter that characterizes the speed of the adsorption process.
(14)
(15)
where qe indicates the equilibrium adsorption capacity (mg/g), qt indicates the adsorption capacity at any time t (min).

Pseudo-second-order kinetics model

Pseudo-second-order (PSO) postulates that the adsorption adheres to a second-order reaction mechanism, signifying that the rate of adsorption occupancy is directly proportional to the square of the number of active sites (Equation (16)) (Duan et al. 2022). Linearized form of the PSO is given by Equation (17). The parameter k2(g/(mg-min)) corresponds to the rate constant of PSO adsorption kinetics.
(16)
(17)
where qt indicates the adsorption capacity (mg/g) at time t (min).

Intra-particle diffusion model

The intra-particle diffusion (IPD) model stands as a widely utilized tool for examining the rate-limiting step in adsorption kinetics. The complex dynamics of adsorption in aqueous solutions encompass various factors, including the mass transfer of HMs (adsorbate), commonly referred to as film diffusion, as well as pore diffusion and surface diffusion (Siraorarnroj et al. 2022). Film diffusion represents an independent step, whereas surface and pore diffusion can occur concurrently. Equation (18) provides the linearized representation of the IPD, offering insights into the intricate mechanisms governing mass transfer and diffusion phenomena during adsorption (Gebrewold et al. 2019). The parameter k3(mg/(g-min1/2)), IPD rate constant, a crucial factor characterizing the rate at which IPD occurs during the adsorption process.
(18)
where C indicates the intercept of intra-particle model plot.

Desorption study

To explore the viability for reuse of PNB and OPNB adsorbents, experiments were in five consecutive adsorption–desorption cycles using 0.1 M HCl as a de-sorbent. The PNB and OPNB were thoroughly washed and dried after every adsorption–desorption cycle. Optimized parameters were selected for carrying out the adsorption–desorption study. Additionally, further desorption experiments with the desorbing agent aimed to determine the optimal duration of the desorption stage, with samples analyzed at intervals of 4, 8, 12, 16, 20, and 24 h. The following equation was used to determine the biochar's desorption efficiency.
(19)
where Cdes indicates the desorbed amount (mg/L) of Cu in the solution, Cads indicates the adsorbed amount of Cu(II) by adsorbents.

Column experiments

Fixed-column experiments were conducted by employing two distinct glass columns, each possessing an internal diameter of 1.8 cm and a height of 50 cm with an effective volume of 101.8 cm3. Figure 1 shows the experimental setup for the fixed-bed column. Both columns were packed with 5 g of adsorbents PNB and OPNB, respectively, with a particle size in the range 500–710 μm. A non-reactive material glass wool, with a thickness of 5 cm (Figure 1), was strategically positioned at the top and bottom of the column to prevent any inadvertent loss of adsorbent and to enhance the stability of the adsorbent bed. To maintain a consistent flow rate of 1 mL/min, a peristaltic pump was employed for the lift of aqueous solution with optimized experimental conditions for both the adsorbents with an initial (Cu(II)) concentration of 125 mg/L. Whereas the pH values were 4.65 in the case of PNB and 6 in case of OPNB. In order to stabilize the flow and eliminate air trapped in the pores, DI water was passed through the column before the experiment began. All experiments were conducted at ambient conditions at a room temperature of 25 ± 2 °C. The 50% breakthrough capacity (Ct/CO = 0.5) was determined using Equation (20).
(20)
Here, Q0.5 indicates the 50% breakthrough capacity (mg/g), t0.5 indicates the breakthrough time at 50% (i.e., at Ce/Ci = 0.5), Ct indicates the effluent metal concentration (mg/L) at time t, Q indicates the flow rate (L/h), m indicates the adsorbent mass (g).
Figure 1

Fixed-bed column reactor.

Figure 1

Fixed-bed column reactor.

Close modal

Thomas model for a fixed-bed study

The Thomas model, which is essentially based on the Langmuir isotherm and second-order kinetics, was employed in the column study to ascertain the dynamic behaviour of fixed-column adsorption. The empirical formula for the Thomas model is presented by the following equation
(21)
The linearized form of the Thomas model is given by Equation (22):
(22)
where KTH indicates the Thomas model constant (L/(mg-h)),

Analytical methods

The determination of the physical properties of PNB and OPNB_3, OPNB_10, OPNB_20, and OPNB_30 such as pore size, SSA and pore volume were meticulously examined using BET analyzer (Quanta chrome, Autosorb iQ3) in inert atmosphere using N2 adsorption at 77 K. The samples were vacuum degassed for 24 hours at 150 °C before BET analysis. The elemental composition of PNB and OPNB was assessed through a CHNSO analyzer (UNICUBE plus) to determine their C, H, N, and O content. The surface morphology images of the PNB, as well as those of the OPNB, were examined using the scanning electron microscope (SEM) system, referred to as EDX-SEM. Fourier transformation infra-red (FTIR) spectroscopy analysis (FTIR L1600312, Agilent Technologies) was conducted to identify the surface functional groups present on the adsorbents (PNB and OPNB) surface. Inductively coupled plasma mass spectrometry (ICP-MS, Agilent 7850) was employed to determine the final concentration of Cu(II) ions present in the SWW.

Characterization of biochar and oxidized biochar

Surface characteristics and elemental properties

Table 2 outlines the surface characteristics and elemental properties of PNB, OPNB_3, OPNB_10, OPNB_20, and OPNB_30 after treatment with 3, 10, 20, and 30% H2O2 solutions, respectively. During the pyrolysis, lignocellulosic biomass underwent decarboxylation and decarbonization, resulted in the production of biochar enriched with surface functional groups, minerals, and ash content. Previous studies reported that H2O2 modification is more effective in removing inorganic material, potentially due to carbon oxidation by H2O2, and making ash removal easier (Bhattacharyya 2016; Wang & Liu 2018).

Table 2

Physicochemical properties of oxidized biochar at different H2O2 concentrations

AdsorbentC (%)H (%)O (%)N (%)O/CH/CSSA (m2/g)Pore volume (cm3/g)Pore size (nm)
PN 48.05 6.51 42.60 0.94 0.177 0.034 1.82 0.003 1.85 
PNB 74.27 2.54 13.14 1.48 0.158 0.018 36.38 0.029 1.98 
3% PNB 73.18 1.49 13.17 1.44 0.180 0.020 135.80 0.129 1.70 
10% PNB 71.97 1.34 15.98 1.46 0.222 0.018 145.75 1.481 1.52 
20% PNB 72.15 1.29 14.69 1.35 0.203 0.017 131.26 0.082 1.90 
30% PNB 72.52 1.29 14.24 1.42 0.196 0.017 68.362 0.029 1.90 
AdsorbentC (%)H (%)O (%)N (%)O/CH/CSSA (m2/g)Pore volume (cm3/g)Pore size (nm)
PN 48.05 6.51 42.60 0.94 0.177 0.034 1.82 0.003 1.85 
PNB 74.27 2.54 13.14 1.48 0.158 0.018 36.38 0.029 1.98 
3% PNB 73.18 1.49 13.17 1.44 0.180 0.020 135.80 0.129 1.70 
10% PNB 71.97 1.34 15.98 1.46 0.222 0.018 145.75 1.481 1.52 
20% PNB 72.15 1.29 14.69 1.35 0.203 0.017 131.26 0.082 1.90 
30% PNB 72.52 1.29 14.24 1.42 0.196 0.017 68.362 0.029 1.90 

In comparison to the initial biomass (PN) with carbon content, PNB and OPNB exhibit increased carbon and nitrogen content, while hydrogen content and the H/C ratio decrease. The oxygen content of the raw material drops after pyrolysis but increases when treated with H2O2. After H2O2 modification, the O content in the biochar rises, attributed to surface carbon oxidation and the proliferation of oxygen-containing functional groups, especially carboxyl groups. The SSA of PNB initially measures around 36.4 m2/g. However, the SSA of OPNB increases when treated with different concentrations of H2O2 (Table 2) reflecting the removal of ash content from the pristine biochar through a DI water washing process following the modification.

Selection of optimized H2O2 concentration for biochar modification

The produced PNB underwent treatment with varying H2O2 concentrations (Table 2) for optimized characteristics like a higher SSA and increased oxygen content. Notably, the SSA of PNB significantly increased to 145.7 m2/g when treated with 10% H2O2, compared to its untreated value of 36.4 m2/g . This substantial increase in SSA was primarily attributed to the removal of organic matter and ash content, subsequently enhancing the porosity of PNB after H2O2 treatment (Wang & Liu 2018; Wongrod et al. 2018, 2019). However, it's noteworthy that the SSA decreased when subjected to higher H2O2 concentrations, such as 20 and 30% (Table 2). This decline in SSA can be attributed to the intensified oxidation process, which disrupted the porous structure and channels within the biochar (Nie et al. 2019). As the optimized properties were obtained at the 10% H2O2 concentration, thereafter all the experiments were conducted with the OPNB_10 and named as OPNB for the further experiments.

FTIR analysis

FTIR spectroscopy was utilized in the present study to offer valuable insights into the structure and confirmation of various molecules. Figure 2 depicts the FTIR spectroscopy of pristine biochar (PNB) and oxidized with 10% H2O2 (OPNB). It is clear from the FTIR spectra that the apparent change after the modification is that the peak of the corresponding group has enlarged (Figure 2). The increased oxygen and carboxyl group levels of the oxidized biochar are consistent with the FTIR results. The peaks at 2,350 cm−1 can be attributed to the CO2 evolution (Ray 2020). The peaks observed at 1,380 cm−1 correspond to O − H bending modes (Ortega et al. 2017). The peak observed at 880 cm−1 is associated with the out-of-plane bending of C − H bonds, a characteristic feature of aromatic rings. This implies the aromatization of the pyrolyzed feedstock (Choudhary et al. 2020). Peaks, such as 1,580 for C = O cm−1 and 1,030 cm−1 for C–O, were detected in the spectrum of PNB and OPNB, signifying abundance of O-containing surface functional groups (Lian et al. 2015; Choudhary et al. 2020). The peaks in the interval of 3,000–3,500 cm−1 are associated with the stretching vibration of –OH in –COOH (Fan et al. 2018; Nie et al. 2019). The peak at 3,220 cm−1 shifted to 3,380 cm−1 after H2O2 treatment suggesting that the shifting and enlargement in peaks takes place after the modification (Bhattacharyya 2016; Wang & Liu 2018).
Figure 2

FTIR spectrum of the adsorbents PNB and OPNB.

Figure 2

FTIR spectrum of the adsorbents PNB and OPNB.

Close modal

Surface morphology

Figure 3 shows the surface morphology of the PNB and OPNB before and after the adsorption process. The surface of PNB appeared rough and randomly ordered heterogeneous, featuring pores of varying sizes. In the case of OPNB a well-developed channelized porous structure was observed (Figure 3). These significant structural changes can be attributed to the removal of inorganic matter and the ash content by the effective loading of H2O2 particles onto the PNB surface.
Figure 3

SEM images before and after adsorption.

Figure 3

SEM images before and after adsorption.

Close modal

Model description and ANOVA

RSM model description

Table 3 presents the results of 30 adsorption experiments, which were designed using CCD under RSM. These results include 8 axial points, 16 cube/factorial points, and 6 duplicates, obtained under the respective experimental conditions. The actual Cu(II) adsorption capacity (Qa) and the predicted adsorption capacity (Qp) for the adsorbents PNB and OPNB presented in Table 3 showed good proximity between the Qa and Qp values.

Table 3

Adsorption capacity (actual and predicted) for the adsorbents PNB and OPNB

Independent variables
Adsorption capacity (mg/g)
PNB
OPNB
S. No.A (mg/L)B (g/L)C (min)DQaQpResidualQaQpResidual
50 0.4 360 25.00 24.88 0.1183 43.25 41.57 1.68 
125 0.4 360 26.36 26.13 0.2329 42.08 41.57 0.5173 
50 0.8 360 21.36 21.13 0.2301 17.85 18.30 −0.4509 
125 0.8 360 21.60 21.69 −0.0848 42.05 41.03 1.02 
50 0.4 960 25.47 26.40 −0.9305 43.99 42.85 1.15 
125 0.4 960 27.96 27.70 0.2540 14.08 14.62 −0.5388 
50 0.8 960 25.68 25.29 0.3967 40.59 40.11 0.4808 
125 0.8 960 25.33 25.90 −0.5664 88.69 89.94 −1.25 
50 0.4 360 25.29 24.59 0.7016 27.86 26.73 1.13 
10 125 0.4 360 26.05 26.71 −0.6607 20.42 20.28 0.1420 
11 50 0.8 360 17.64 18.16 −0.5180 88.95 87.79 1.15 
12 125 0.8 360 20.66 19.59 1.07 40.41 41.57 −1.16 
13 50 0.4 960 30.94 31.12 −0.1792 38.63 37.50 1.13 
14 125 0.4 960 33.20 33.29 −0.0949 91.57 91.61 −0.0398 
15 50 0.8 960 27.22 27.32 −0.0978 38.09 41.57 −3.48 
16 125 0.8 960 28.42 28.80 −0.3823 38.66 40.26 −1.60 
17 12.5 0.6 660 4.5 25.11 25.03 0.0750 49.06 48.04 1.02 
18 162.5 0.6 660 4.5 27.82 27.76 0.0539 19.53 21.08 −1.54 
19 87.5 0.2 660 4.5 31.79 31.57 0.2149 85.55 84.58 0.9663 
20 87.5 660 4.5 23.25 23.34 −0.0860 29.03 29.43 −0.4002 
21 87.5 0.6 60 4.5 18.86 19.47 −0.6069 87.29 89.13 −1.84 
22 87.5 0.6 1,260 4.5 30.94 30.20 0.7358 85.74 84.73 1.01 
23 87.5 0.6 660 1.5 22.29 22.18 0.1104 43.91 41.57 2.34 
24 87.5 0.6 660 7.5 24.81 24.79 0.0185 6.57 5.30 1.27 
25 87.5 0.6 660 4.5 28.20 29.32 −1.12 44.06 45.36 −1.30 
26 87.5 0.6 660 4.5 29.15 29.32 −0.1725 31.26 32.16 −0.8988 
27 87.5 0.6 660 4.5 29.55 29.32 0.2260 43.06 42.74 0.3202 
28 87.5 0.6 660 4.5 29.62 29.32 0.3015 41.67 41.57 0.0993 
29 87.5 0.6 660 4.5 29.86 29.32 0.5366 50.42 49.17 1.25 
30 87.5 0.6 660 4.5 29.55 29.32 0.2253 40.37 42.55 −2.18 
Independent variables
Adsorption capacity (mg/g)
PNB
OPNB
S. No.A (mg/L)B (g/L)C (min)DQaQpResidualQaQpResidual
50 0.4 360 25.00 24.88 0.1183 43.25 41.57 1.68 
125 0.4 360 26.36 26.13 0.2329 42.08 41.57 0.5173 
50 0.8 360 21.36 21.13 0.2301 17.85 18.30 −0.4509 
125 0.8 360 21.60 21.69 −0.0848 42.05 41.03 1.02 
50 0.4 960 25.47 26.40 −0.9305 43.99 42.85 1.15 
125 0.4 960 27.96 27.70 0.2540 14.08 14.62 −0.5388 
50 0.8 960 25.68 25.29 0.3967 40.59 40.11 0.4808 
125 0.8 960 25.33 25.90 −0.5664 88.69 89.94 −1.25 
50 0.4 360 25.29 24.59 0.7016 27.86 26.73 1.13 
10 125 0.4 360 26.05 26.71 −0.6607 20.42 20.28 0.1420 
11 50 0.8 360 17.64 18.16 −0.5180 88.95 87.79 1.15 
12 125 0.8 360 20.66 19.59 1.07 40.41 41.57 −1.16 
13 50 0.4 960 30.94 31.12 −0.1792 38.63 37.50 1.13 
14 125 0.4 960 33.20 33.29 −0.0949 91.57 91.61 −0.0398 
15 50 0.8 960 27.22 27.32 −0.0978 38.09 41.57 −3.48 
16 125 0.8 960 28.42 28.80 −0.3823 38.66 40.26 −1.60 
17 12.5 0.6 660 4.5 25.11 25.03 0.0750 49.06 48.04 1.02 
18 162.5 0.6 660 4.5 27.82 27.76 0.0539 19.53 21.08 −1.54 
19 87.5 0.2 660 4.5 31.79 31.57 0.2149 85.55 84.58 0.9663 
20 87.5 660 4.5 23.25 23.34 −0.0860 29.03 29.43 −0.4002 
21 87.5 0.6 60 4.5 18.86 19.47 −0.6069 87.29 89.13 −1.84 
22 87.5 0.6 1,260 4.5 30.94 30.20 0.7358 85.74 84.73 1.01 
23 87.5 0.6 660 1.5 22.29 22.18 0.1104 43.91 41.57 2.34 
24 87.5 0.6 660 7.5 24.81 24.79 0.0185 6.57 5.30 1.27 
25 87.5 0.6 660 4.5 28.20 29.32 −1.12 44.06 45.36 −1.30 
26 87.5 0.6 660 4.5 29.15 29.32 −0.1725 31.26 32.16 −0.8988 
27 87.5 0.6 660 4.5 29.55 29.32 0.2260 43.06 42.74 0.3202 
28 87.5 0.6 660 4.5 29.62 29.32 0.3015 41.67 41.57 0.0993 
29 87.5 0.6 660 4.5 29.86 29.32 0.5366 50.42 49.17 1.25 
30 87.5 0.6 660 4.5 29.55 29.32 0.2253 40.37 42.55 −2.18 

Equations (23) and (24) represent the response (adsorption capacity) of the CCD model in the variable form for the adsorbents PNB and OPNB, respectively. These equations depend on the different independent factors.
(23)
(24)

Values with positive sign in Equations (23) and (24) signify an incremental effect of the variable on the adsorption capacity, while values with negative sign indicate a decremental effect of the variable on the Cu(II) adsorption capacity for PNB and OPNB, respectively.

The predicted adsorption capacity values from Equations (23) and (24) exhibit a close agreement with the experimental values, as presented in Table 3. Assessment of the values presented in Table 3 reveals minimal residuals between the Qa and Qp. Model suitability was assessed through correlation coefficients R2 and adj. R2. Referring to the analysis of variance (ANOVA) results (Tables 4 and 5), the R2 and adj. R2 values were determined to be 0.96 and 0.92, respectively, in case of PNB. Whereas in case of OPNB these values were found to be 0.99 and 0.98. The high R2 values underscore a strong correlation between the predicted and actual response values. Additionally, adj. R2 value indicates that models describe the 92 and 98% variation in Cu(II) adsorption is accounted for the independent variables in case of PNB and OPNB, respectively. While only 8% for PNB and 2% for OPNB, indicating the portion of variability not captured by these models.

Table 4

Analysis of variance for the copper adsorption using PNB

SourceSum of squaresdfMean squareF-valuep-valueR2Adj. R2
Model 423.30 14 30.24 66.61 <0.0001 significant 0.9694 0.9265 
11.18 11.18 24.63 0.0002    
101.74 101.74 224.14 <0.0001    
172.96 172.96 381.05 <0.0001    
10.26 10.26 22.61 0.0003    
AB 0.4738 0.4738 1.04 0.3231    
AC 0.0029 0.0029 0.0065 0.9368    
AD 0.7632 0.7632 1.68 0.2143    
BC 6.92 6.92 15.24 0.0014    
BD 7.18 7.18 15.82 0.0012    
CD 25.05 25.05 55.18 <0.0001    
A² 14.64 14.64 32.26 <0.0001    
B² 5.98 5.98 13.17 0.0025    
C² 34.49 34.49 75.98 <0.0001    
D² 58.41 58.41 128.68 <0.0001    
Residual 6.81 15 0.4539      
Lack of Fit 5.05 10 0.5051 1.44 0.3616 not significant   
Pure Error 1.76 0.3516      
Cor Total 430.11 29       
SourceSum of squaresdfMean squareF-valuep-valueR2Adj. R2
Model 423.30 14 30.24 66.61 <0.0001 significant 0.9694 0.9265 
11.18 11.18 24.63 0.0002    
101.74 101.74 224.14 <0.0001    
172.96 172.96 381.05 <0.0001    
10.26 10.26 22.61 0.0003    
AB 0.4738 0.4738 1.04 0.3231    
AC 0.0029 0.0029 0.0065 0.9368    
AD 0.7632 0.7632 1.68 0.2143    
BC 6.92 6.92 15.24 0.0014    
BD 7.18 7.18 15.82 0.0012    
CD 25.05 25.05 55.18 <0.0001    
A² 14.64 14.64 32.26 <0.0001    
B² 5.98 5.98 13.17 0.0025    
C² 34.49 34.49 75.98 <0.0001    
D² 58.41 58.41 128.68 <0.0001    
Residual 6.81 15 0.4539      
Lack of Fit 5.05 10 0.5051 1.44 0.3616 not significant   
Pure Error 1.76 0.3516      
Cor Total 430.11 29       
Table 5

Analysis of variance for the copper adsorption using OPNB

SourceSum of squaresdfMean squareF-valuep-valueR2Adj. R2
Model 16,463.20 14 1,175.94 335.54 <0.0001 significant 0.9938 0.9874 
9,462.78 9,462.78 2,700.04 <0.0001    
5,593.39 5,593.39 1,595.98 <0.0001    
42.91 42.91 12.24 0.0032    
99.44 99.44 28.37 <0.0001    
AB 740.96 740.96 211.42 <0.0001    
AC 3.31 3.31 0.9439 0.3467    
AD 29.86 29.86 8.52 0.0106    
BC 0.9006 0.9006 0.2570 0.6196    
BD 26.70 26.70 7.62 0.0146    
CD 8.28 8.28 2.36 0.1450    
A² 20.39 20.39 5.82 0.0291    
B² 422.24 422.24 120.48 <0.0001    
C² 3.33 3.33 0.9512 0.3449    
D² 21.36 21.36 6.09 0.0261    
Residual 52.57 15 3.50      
Lack of Fit 30.54 10 3.05 0.6932 0.7097 not significant   
Pure Error 22.03 4.41      
Cor Total 16,515.77 29       
SourceSum of squaresdfMean squareF-valuep-valueR2Adj. R2
Model 16,463.20 14 1,175.94 335.54 <0.0001 significant 0.9938 0.9874 
9,462.78 9,462.78 2,700.04 <0.0001    
5,593.39 5,593.39 1,595.98 <0.0001    
42.91 42.91 12.24 0.0032    
99.44 99.44 28.37 <0.0001    
AB 740.96 740.96 211.42 <0.0001    
AC 3.31 3.31 0.9439 0.3467    
AD 29.86 29.86 8.52 0.0106    
BC 0.9006 0.9006 0.2570 0.6196    
BD 26.70 26.70 7.62 0.0146    
CD 8.28 8.28 2.36 0.1450    
A² 20.39 20.39 5.82 0.0291    
B² 422.24 422.24 120.48 <0.0001    
C² 3.33 3.33 0.9512 0.3449    
D² 21.36 21.36 6.09 0.0261    
Residual 52.57 15 3.50      
Lack of Fit 30.54 10 3.05 0.6932 0.7097 not significant   
Pure Error 22.03 4.41      
Cor Total 16,515.77 29       

Analysis of variance

Moreover, the ANOVA for Cu(II) adsorption was used during the study for determining the adequacy of the model. Tables 4 and 5 presented ANOVA for PNB and OPNB adsorbents, respectively, where the significance of each variable was determined through the Fischer's test coefficients (F) and probability (P) values. Assessing a regression models significance involves examining the F-value for overall effectiveness and the P value for individual coefficients significance. Higher F values indicate stronger relationships, while lower P values signal greater confidence in specific variable contributions (Lan et al. 2011). This dual evaluation enhances our understanding of model validity and variable impact. In Tables 4 and 5, it is evident that the RSM-proposed model is highly significant, as indicated by a high F-value of 66.61 and a minimal P value of less than 0.0001 in the case of PNB. Meanwhile these values were found to be 335.54 and <0.0001 in the case of OPNB. The individual coefficients, including linear, square, and quadratic terms, also exhibit high significance, with large F values and correspondingly low P values (Tables 4 and 5). In the case of PNB, terms like AB, AC, and AD are deemed non-significant, as indicated by P > 0.05. Similarly, for OPNB, terms such as AC, BC, CD, and C² are considered non-significant with P > 0.05. Furthermore, P-value for PNB 0.3616 and 0.7079 for OPNB suggests that the lack-of-fit has no significance, affirming the validity of the regression analysis for Cu(II) adsorption.

Three-dimensional response surface models

The three-dimensional (3D) surface plots depicting adsorption uptake (Figures 4 and 5) serve as graphical representations of the quadratic equation. The graphical representation provides a comprehensive visual analysis of the interaction between adsorption capacity and different variables throughout the adsorption process. Figures 4(a), 4(d), 4(e), 5(a), 5(d) and 5(e) demonstrate that increasing the concentration of Cu(II) ions in the range of 50–125 mg/L increases the adsorption capacity of both adsorbents. Moreover, beyond a certain limit, elevations in Cu(II) concentration resulted in only a marginal reduction in Cu(II) levels, indicating the saturation of adsorbent sites at a constrained concentration (Figure 4(a) and 4(b)). Notably, in the case of OPNB, the adsorption capacity for Cu(II) surpassed that of PNB. This can be ascribed to the increased adsorption sites (high SSA) of OPNB, necessitating a higher concentration of Cu(II) for saturation.
Figure 4

Three-dimensional response surfaces depicting adsorption capacity in relation to (a) copper concentration and contact time, (b) PNB dose and time, (c) copper concentration and time, (d) time and pH, (e) copper concentration and PNB dose, (f) copper concentration and pH, and (g) PNB dose and pH.

Figure 4

Three-dimensional response surfaces depicting adsorption capacity in relation to (a) copper concentration and contact time, (b) PNB dose and time, (c) copper concentration and time, (d) time and pH, (e) copper concentration and PNB dose, (f) copper concentration and pH, and (g) PNB dose and pH.

Close modal
Figure 5

Three-dimensional response surfaces depicting adsorption capacity in relation to (a) copper concentration and time, (b) OPNB dose and time, (c) copper concentration and contact time, (d) time and pH, (e) copper concentration and OPNB dose, (f) copper concentration and pH, and (g) OPNB dose and pH.

Figure 5

Three-dimensional response surfaces depicting adsorption capacity in relation to (a) copper concentration and time, (b) OPNB dose and time, (c) copper concentration and contact time, (d) time and pH, (e) copper concentration and OPNB dose, (f) copper concentration and pH, and (g) OPNB dose and pH.

Close modal

Figure 4(b), 4(d) and 4(f) illustrate the effect of adsorbent (PNB) dosage on the adsorption of Cu(II). Likewise, Figures 5(b), 5(d), and 5(f) depict the impact of OPNB dosage on Cu(II) adsorption of Cu(II). Figure 4(f) for PNB and Figure 4(f) for OPNB show the adsorption capacity as pH and adsorbent dosages vary. Also, the Figures 4(b), 4(d), 5(b) and 5(d) illustrate a decrease in Cu(II) uptake as the adsorbent dosage increases from 0.4 to 1.0 g/L for both the adsorbents. This paradox can be ascribed to the fact that, despite the increase in adsorption sites at higher adsorbent doses, the adsorption reaction does not saturate, leading to a decline in Cu(II) uptake. Moreover, the cluster of the adsorbents at higher doses may contribute to the reduction in Cu(II) removal capacity. Interestingly, optimal Cu(II) adsorption occurs at lower adsorbent dosages and higher pH levels. This occurrence is likely elucidated by the heightened accessibility of adsorption sites at higher pH values, enhancing Cu(II) interaction with the adsorbents. This phenomenon underscores the characteristic of the chelation mechanism, which can be ascribed to the reduced competitive adsorption of H+ at higher pH (Huang & Chen 2009; Tan et al. 2012). Nevertheless, with an increasing quantity of adsorbent, there is a rise in effective surface area and the ratio of adsorbent to adsorbate, leading to a decrease in the uptake of Cu(II) (Ahmad & Hasan 2016).

In Figure 4(a)–4(c), the results distinctly demonstrate an initial increase in adsorption capacity with contact time for PNB, reaching a peak, and then experiencing a slight decrease. Likewise, for the adsorbent OPNB, the adsorption capacity exhibits an increase over time, followed by a decrease, as depicted in Figure 5(a)–5(c). The increase in adsorption capacity at initial stage with time can be due to availability of more vacant sites for the adsorption. This can relate with the enhanced SSA and increased oxygen content (Table 2) resulting from the oxidation of biochar. Consequently, the elevated concentration of O-containing functional groups contributed to the increment in Cu(II) uptake. Nevertheless, a subsequent increase in contact time resulted in a reduction in adsorption capacity, attributed to the desorption of Cu(II) into the solution after attaining the equilibrium.

In the complicated realm of HM removal, the pH of the solution contributes as an important factor (Ahmad & Hasan 2016). Influence of pH is illustrated in the Figure 4(c), 4(e) and 4(f) for the adsorption capacity of PNB. Similarly Figure 5(c), 5(e) and 5(f) demonstrate the influence of pH on Cu(II) adsorption for the adsorbent OPNB. A meticulous examination of the data in Figures 4(c) and 5(c) elucidates a nuanced relationship between the adsorption rate of Cu(II) and the pH variation, revealing a notable augmentation as pH values ascend from 3 to 6. The pivotal role of pH becomes distinct below the pH threshold of 3.0, where a competitive interaction occurs between H+ ions and Cu(II) ions for the adsorbent surface. This competition introduces a barrier to the approach of Cu(II), dictated by repulsion forces, resulting in a diminished metal removal rate (Luef et al. 1991). This interaction implies that, at pH levels below 3.0, protons compete with Cu(II) for ligand binding sites and complex formation, which contributes to the observed decrease in metal removal. As pH ascends, H+ ions decrease, increases Cu(II) adsorption. This observed increase in adsorption capacity with an increasing pH can be elucidated by the heightened negative charge characterizing the surface of the adsorbent (Kalavathy et al. 2009).

Optimization of the process parameters using desirability function

The final step in the RSM involves determining the optimal conditions for achieving maximum Cu(II) uptake yields through the adsorption process using PNB and OPNB as adsorbents. For optimization purposes, the software integrates global information from the model, considering response surfaces, and the interactions among all the process variables to identify the optimum conditions for the overall process. The evaluation of optimal conditions is performed using the desirability function (d). To ensure comprehensive coverage of parameter variations, the lower limit (d = 0, representing the response at its undesirable limit) and upper limit (d = 1.0, signifying a highly desirable limit of the response) were set.

To achieve optimal conditions for Cu(II) uptake through the adsorption process, the adsorbent initial Cu(II) concentration was set at the maximum level, adsorbent dosages were set at a minimum level, while contact time was set at minimum and pH at a maximum level for maximum desirability. Figures 5 and 6 depicts the optimization plot of different variables and the response generated from optimum operating variables through numerical optimization. Subsequently, the statistical software provided the optimum conditions. For both adsorbents, the ideal Cu(II) uptake parameters were determined to be an initial (Cu(II)) concentration of 125 mg/L, an adsorbent dosage of 0.4 g/L, and a contact time of 360 min. While the pH values were 4.65 in case of PNB and 6 in case of OPNB (Figure 6). These optimal conditions resulted in Cu(II) adsorption of 27.89 mg/g for PNB (Figure 6) with a desirability of 0.90. Similarly, these values were 91.61 mg/g and 1.0 in case of OPNB (Figure 7). Furthermore, to verify the adequacy and validity of the optimization procedure, experiments were replicated in triplicate under the optimized conditions, and the outcomes were compared with these predicted response values. Subsequently, results from confirmation experiments revealed adsorption capacities of 26.34 ± 0.48 and 90.42 ± 0.34, for PNB and OPNB, respectively, and closely aligning with the predicted values. This consistency affirms the suitability of the developed regression model. It is crucial to notice that these optimal values remain valid within the specified range of Cu(II) adsorption process variables. Consequently, the validity and adequacy of the models have been verified. Furthermore, the optimal values obtained from this process were used to carry out the isotherm and kinetics study.
Figure 6

Desirability of Cu(II) adsorption for the adsorbent PNB.

Figure 6

Desirability of Cu(II) adsorption for the adsorbent PNB.

Close modal
Figure 7

Desirability of Cu(II) adsorption for the adsorbent OPNB.

Figure 7

Desirability of Cu(II) adsorption for the adsorbent OPNB.

Close modal

Isotherm models

Figure 8 depicts the Langmuir, Freundlich and Temkin isotherm models with their correlation coefficient based on the experimental data. The analysis of the Langmuir isotherm for Cu(II) adsorption using PNB and OPNB revealed a strong fit to the experimental data, as evidenced by high R2 of 0.99 for both the adsorbents (Figure 8(a)). For the Cu(II) adsorption, the experimental value for the Langmuir model was determined to be 26.58 mg for PNB and 91.54 mg for OPNB. However, qm for PNB and OPNB were found to be 29.49 and 102.04 mg/g, respectively, using the Langmuir model. (Table 6). Furthermore, the RL equilibrium parameter, indicating favourable adsorption when 0 < RL < 1, was calculated to be 0.087–0.188 for PNB and 0.075–0.167 for OPNB, confirming the favourable adsorption on the surfaces of the adsorbents.
Table 6

Isotherm models parameters for the adsorption of copper

MetalAdsorbentQexp.Langmuir model
Freundlich model
Temkin model
Qmax (mg/g)KL (L/mg)R2KF [L/(g*n)]nR2BT (kJ/mol)KTR2
Cu PNB 26.58 29.49 0.051 0.99 8.35 4.46 0.88 1.706 4.669 0.81 
OPNB 91.54 102.04 0.061 0.99 24.04 3.44 0.95 −1.877 19.774 0.91 
MetalAdsorbentQexp.Langmuir model
Freundlich model
Temkin model
Qmax (mg/g)KL (L/mg)R2KF [L/(g*n)]nR2BT (kJ/mol)KTR2
Cu PNB 26.58 29.49 0.051 0.99 8.35 4.46 0.88 1.706 4.669 0.81 
OPNB 91.54 102.04 0.061 0.99 24.04 3.44 0.95 −1.877 19.774 0.91 
Figure 8

Adsorption isotherms models: (a) Langmuir, (b) Freundlich, and (c) Temkin.

Figure 8

Adsorption isotherms models: (a) Langmuir, (b) Freundlich, and (c) Temkin.

Close modal

The Freundlich isotherm (Figure 8(b)) model exhibited R2 values were determined to be 0.88 and 0.95 for PNB and OPNB, respectively, indicating a suitable-fitted model for Cu(II) adsorption (Figure 8(b)). The Freundlich constants (KF), reflecting the energy of adsorption, were 8.35 mg/g for PNB and 24.04 mg/g for OPNB. The intensity of adsorption (n) suggested cooperative adsorption for PNB (4.46) and normal adsorption for OPNB (3.44).

Qexp – Experimental adsorption capacity

Also, the Temkin model displayed a R2 of 0.81 for PNB and 0.91 for OPNB (Figure 8(c)), along with corresponding adsorption constants BT and KT presented in Table 6. It was indicated by the lower R2 values for PNB and OPNB that the Temkin model did not adequately fit the data (Table 6). These findings indicated a lower degree of surface heterogeneity and compatibility with both the Langmuir and Freundlich isotherms.

As here in the present study of value of R2 obtained for Temkin model and Freundlich model was found to be lower than those obtained in the Langmuir model. High R2 (0.99) value and closeness among the experimental and predicted values from models of adsorption capacity suggested that the monolayer adsorption phenomenon for the uptake of Cu(II) for using the absorbents PNB and OPNB.

Also, in most AMD the discharged effluent typically encompasses Cu(II) concentrations within the interval of 3–138 mg/L (Edraki et al. 2005; Ayora et al. 2016). The present study showed that the OPNB has an adsorption capacity of up to 102.04 mg/g and it can efficiently remove Cu(II) up to 125 mg/L. Thus, the selected concentration holds practical significance for the treatment of AMD. During this study, it was found that the OPNB can efficiently remove Cu(II) up to 125 mg/L with an adsorption capacity of up to 102.04 mg/g. Thus, the chosen concentration holds practical significance for AMD treatment.

Kinetics study

In general, various kinetic models are employed to understand and control the mechanisms of adsorption processes, encompassing aspects like mass transfer and chemical reactions. The rate of adsorption, a vital measure of adsorbent efficiency, can offer insights into the underlying adsorption mechanisms. In the present study the adsorption kinetics of Cu(II) uptake were examined using three models, namely PFO, PSO and IPD. Figure 9 shows the three-adsorption kinetics model namely PFO, PSO and IPD. Table 7 summarizes the important parameters that were found from the kinetic models.
Table 7

Kinetics study parameters for different models

MetalAdsorbentqexp. (mg/g)Pseudo-first-order
Pseudo-second-order
Intra-particle diffusion
qe (mg/g)k1R2qe (mg/g)k2R2Ik3 (mg/g-min1/2)R2
Cu PNB 26.64 63.65 0.0117 0.90 30.30 0.00052 0.98 −2.14 1.48 0.81 
OPNB 91.13 234.96 0.0105 0.97 116.27 0.00006 0.98 8.30 4.20 0.89 
MetalAdsorbentqexp. (mg/g)Pseudo-first-order
Pseudo-second-order
Intra-particle diffusion
qe (mg/g)k1R2qe (mg/g)k2R2Ik3 (mg/g-min1/2)R2
Cu PNB 26.64 63.65 0.0117 0.90 30.30 0.00052 0.98 −2.14 1.48 0.81 
OPNB 91.13 234.96 0.0105 0.97 116.27 0.00006 0.98 8.30 4.20 0.89 
Figure 9

Adsorption kinetics models: (a) PFO, (b) PSO, and (c) IPD model for Cu(II) adsorption.

Figure 9

Adsorption kinetics models: (a) PFO, (b) PSO, and (c) IPD model for Cu(II) adsorption.

Close modal

In Table 7, constants k1, k2 and k3 represent the rate constants for the adsorption kinetics of PFO, PSO and IPD, respectively. R2 serves as the correlation coefficient, gauging the degree of agreement between model-predicted values and experimental data. To assess the appropriateness of each model, all of them were subjected to linear regression analysis with the results of experimental data. The R2 and the agreement between Qexp. and predicted Qp values serve as criteria for model applicability.

The results in Table 7 demonstrate that the R2 values of the PSO model were notably high, exceeding 0.98 for both the adsorbents PNB and OPNB. Furthermore, the predicted values closely corresponded to the experimental values obtained under identical conditions. The strong correlation between the adsorption process of Cu(II) and thus the PSO principles suggests that the adsorption rate is directly proportional to the square of the number of available adsorption sites (Duan et al. 2022). These findings imply that the adsorption of Cu(II) onto both PNB and OPNB involved a combination of physisorption and chemisorption, with chemisorption dominating the overall adsorption process. The PFO and IPD models, on the other hand, showed lower correlation coefficients than PSO and were found to be insufficient in explaining Cu(II) adsorption onto biochar as well as oxidized biochar.

Desorption study

The regeneration of PNB and OPNB was assessed through five adsorption–desorption cycles using 0.1 M HCl as a desorbing agent. The ideal desorption time was found to be 12 h in subsequent experiments using 0.1 M HCl as the desorbing agent, as illustrated in Figure 10. Following this, adsorption–desorption was conducted for five cycles with the use of 12 h of desorption time and 0. 1 M HCl as a de-adsorbent to determine their reusability. As depicted in the Figure 10, the adsorption decreased to roughly 50% in three consecutive cycles for the PNB. On the other hand, it was above 80% of in the fourth cycle. These results indicate that the OPNB has the better reusability than that of PNB. This can be attributed to the fact that OPNB have more pores which are reversible in nature after the desorption with the 0.1 M HCl. This suggests that the oxidation of the biochar helps in significantly improving the porosity and hence can be utilized even up to the fourth cycle with only a slight decrease in adsorption capacity of just 19% even after the multiple adsorption–desorption cycles.
Figure 10

Variation of adsorption capacity for the adsorbents PNB and OPNB.

Figure 10

Variation of adsorption capacity for the adsorbents PNB and OPNB.

Close modal

Column study

Figure 11(a) illustrates the breakthrough curves of the adsorbents PNB and OPNB for the adsorption of Cu(II). Values for the Thomas constants were calculated using the linearized model (Equation (22)) and are presented in Table 8. Breakthrough initiation occurred after 6 hours of column operation in case of the PNB adsorbent achieving 50% breakthrough in about 14 h (Figure 11(a)). Whereas in the case of OPNB breakthrough initiation takes place after 10 h and the breakthrough is achieved after 21 h. Bed exhaustion was successfully accomplished in approximately 48 h for both the adsorbents. In the first adsorption cycles, the 50% breakthrough capacity was calculated using Equation (20), and the values obtained were 21.0 mg/g for PNB and 31.7 mg/g for OPNB, with an initial Cu(II) concentration of 125 mg/L. These values are in line with the values determined using the Thomas model (Table 8). Afterwards, both columns were desorbed and used for the second breakthrough using a 0.1 M HCl desorbing solution. During the second adsorption cycle the beginning of breakthrough was observed after 6 h of column operation in the case of PNB, as shown in Figure 11(b), however, the adsorption capacity reduced to half after the second cycle. After operating the column for approximately 24 h, complete bed exhaustion was noted in the case of PNB. On the other hand, the breakthrough initiated after 10 h during the second cycle also, with the exhaustion of the bed after the 24 h. These findings demonstrated that the adsorption capacity of PNB reduces in the second cycle, while the OPNB retains its adsorption capacity even after the second adsorption cycle. The OPNB column showed more adsorption capacity than the PNB even after the second adsorption cycle. This demonstrates that the oxidation of biochar enhances the properties which helped in improving the adsorption capacity and also the reusability of the modified biochar.
Table 8

Constants of the column study

AdsorbentsInitial Cu(II) concentration C0 (mg/L)Adsorbent mass, m (g)Flow rate, Q (L/h)KTH (L/(mg-h))qt (mg/g)R2
PNB 125 0.06 0.0017 22.48 0.94 
OPNB 125 0.06 0.0034 31.70 0.94 
AdsorbentsInitial Cu(II) concentration C0 (mg/L)Adsorbent mass, m (g)Flow rate, Q (L/h)KTH (L/(mg-h))qt (mg/g)R2
PNB 125 0.06 0.0017 22.48 0.94 
OPNB 125 0.06 0.0034 31.70 0.94 
Figure 11

Graphs for (a) first breakthrough curve, (b) second breakthrough curve, and (c) Thomas model.

Figure 11

Graphs for (a) first breakthrough curve, (b) second breakthrough curve, and (c) Thomas model.

Close modal

The Cu(II) adsorption capacity in the fixed-bed column (31.7 mg/g) exhibited a lower value compared to the batch sorption capacity (90.88 mg/g). This disparity can be attributed to the existence of a broad mass transfer zone, ultimately leading to a reduction in the effective bed capacity (Choudhary et al. 2020).

Mechanism of HM removal

The expected mechanism for copper removal can include ion exchange, physical adsorption, complexation, electrostatic attraction and precipitation. Out of these mechanisms precipitation can be ruled out since copper precipitation takes place beyond pH 9 while for the experiments the pH range maintained was 6–7. The adsorption capacity of the H2O2-oxidized biochar (OPNB) increased to about 3.5 times compared to the parent biochar. This can be attributed to the physical adsorption, electrostatic attraction, ion exchange and complexation.

The SSA of OPNB is significantly higher at 145.7 m2/g compared to 36.4 m2/g for PNB (Table 2). The increased SSA of OPNB offers more adsorption sites for Cu(II), surpassing PNB and hence demonstrating higher adsorption capacity. Also, the micro-and mesopores within the biochar trapped the Cu(II). Moreover, within the pH range (6–7) under consideration, the intricate process of Cu(II) adsorption onto the surfaces of PNB and OPNB was predominantly governed by physical adsorption through the confluence of fundamental forces. These encompass the van der Waals forces, which involve the attractive forces between molecules, the London forces, arising from temporary dipoles, and electrostatic forces such as dipole–dipole attractions (Bashir et al. 2022; Osman et al. 2023). The interplay of these forces contributes to the meticulous and dynamic nature of the adsorption phenomenon, highlighting the complex interactions that dictate the affinity of Cu(II) ions for the biochar surfaces. Hence, these weak forces signify physisorption as the predominant adsorption mechanism, characterized by its reversible nature.

FTIR spectra analysis of PNB and the OPNB undertaken in this study, as illustrated in Figure 2, revealed the presence of essential functional groups such as –COOH and –OH on the biochar surfaces. These functional groups aid in adsorption through the electrostatic attraction. Experimental determination of pHPZC values of 5.7 for PNB and 5.1 for OPNB, demonstrating that the biochar surfaces maintained a negative charge within the pH range of 6–7 throughout the experiments, as illustrated in Equation (25) (Ofudje et al. 2015). This outlines the reciprocation between HM ions (M2+) in solution and carboxyl groups on the biochar surface (Swiatkowski et al. 2004).
(25)

In this process (Equation (25)), the O–H bond is disrupted, and new bonds are formed between O and M2+ once the metal ions migrate to the biochars (PNB and OPNB) surface (Wang & Liu 2018). The H2O2 modification increased the concentration of carboxyl groups in PNB, contributing substantially to the enhanced Cu(II) adsorption capacity of the biochar (Wang & Liu 2018). Furthermore, the pHPZC of OPNB is 5.1 which is lower than the pHPZC of PNB 5.7. Hence, the surface of OPNB is expected to be more negatively charged as that of PNB, allowing higher Cu(II) adsorption on OPNB surface.

Furthermore, the biochar surfaces contain exchangeable cations such as calcium, sodium etc. as reported elsewhere (Bashir et al. 2022). These cations replaced by Cu(II) ions through an ion exchange process which facilitated the swapping of metal ions for the native cations on the PNB and OPNB surface. Thus, this swapping/interchange of ions helps in removing copper metal from the aqueous solution.

Comparison with other studies

The comparative analysis of Cu(II) adsorption capacities employing various adsorbents is outlined in Table 9. The findings indicate that the adsorption capacity achieved with OPNB was 102.4 mg/g. Notably, Table 9 illustrates a significantly higher adsorption capacity for OPNB compared to its precursor biochar (PNB). Additionally, the table reveals that the modification of PNB through H2O2 treatment resulted in a comparatively higher adsorption capacity than that of the other adsorbents (Table 9). Thus, it can be interpretated that the oxidation of the biochar increases the adsorption capacity much higher than the original biochar.

Table 9

Comparative analysis of Cu(II) adsorption capacity using OPNB with the literature

S. no.AdsorbentModifying agentExperimental conditionsHeavy metalC (%)O (%)SSA (m2/g)Q (mg/g)Ref.
Brewers draff KOH Cu(II) concentration 0.2–2 mM, time 1–48 h, biochar/activated biochar dose 0.2 g/L Cu(II) 69.1 NR 11.6 10.3 Trakal et al. (2014)  
Cymbopogon schoenanthus L. Spreng H2O2 Cu(II) concentration 10 mg/L, time 1–48 h, biochar/activated biochar dose 0.1–5 g/L, pH 2–8 Cu(II) NR NR 27.3 53.8 Bhattacharyya (2016)  
Yak manure H2O2 Cu(II) concentration 0–200 mg/L, time 1–24 h, biochar/modified biochar dose 0.1 g/L, pH 5.5 Cu(II) 40.0 29.1 6.3 64.9 Wang & Liu (2018)  
Bamboo H2O2 Cu(II) concentration 2.5–25 mg/L, time biochar/modified, pH 5–7 Cu(II) 81.8 15.0 207.3 2.1 Nie et al. (2019)  
OPNB H2O2 Cu(II) concentration 12.5–1,625 mg/L, time 1–21 h, biochar/modified 0.2–1 g/L, pH 1.5–7.5 Cu(II) 71.9 15.9 145.7 102.0 Present study 
S. no.AdsorbentModifying agentExperimental conditionsHeavy metalC (%)O (%)SSA (m2/g)Q (mg/g)Ref.
Brewers draff KOH Cu(II) concentration 0.2–2 mM, time 1–48 h, biochar/activated biochar dose 0.2 g/L Cu(II) 69.1 NR 11.6 10.3 Trakal et al. (2014)  
Cymbopogon schoenanthus L. Spreng H2O2 Cu(II) concentration 10 mg/L, time 1–48 h, biochar/activated biochar dose 0.1–5 g/L, pH 2–8 Cu(II) NR NR 27.3 53.8 Bhattacharyya (2016)  
Yak manure H2O2 Cu(II) concentration 0–200 mg/L, time 1–24 h, biochar/modified biochar dose 0.1 g/L, pH 5.5 Cu(II) 40.0 29.1 6.3 64.9 Wang & Liu (2018)  
Bamboo H2O2 Cu(II) concentration 2.5–25 mg/L, time biochar/modified, pH 5–7 Cu(II) 81.8 15.0 207.3 2.1 Nie et al. (2019)  
OPNB H2O2 Cu(II) concentration 12.5–1,625 mg/L, time 1–21 h, biochar/modified 0.2–1 g/L, pH 1.5–7.5 Cu(II) 71.9 15.9 145.7 102.0 Present study 

Oxidation of PNB significantly enhances its physicochemical properties, including porous structure, functional groups, and textural characteristics. RSM was utilized to design and optimize the sorption of Cu(II) ions for PNB and OPNB. The optimal values obtained were 125 mg/L initial Cu(II) concentration, 0.4 g/L adsorbent dosage, 360 min contact time and a pH of 6 with the desirability of 1 for OPNB. These optimal conditions resulted in Cu(II) adsorption of 102.04 mg/g in case of OPNB about 3.46 times more than that of PNB. Langmuir isotherm and the PSO model accurately fit the experimental results. Desorption study revealed that the OPNB has better reusability up to four cycles without losing its adsorption capacity due to its regeneration using the 0.1 M HCl. On the other hand, results of the column study demonstrated the superior performance of biochar after the oxidation using the H2O2. The superior performance of OPNB was attributed to its higher surface area, well-developed porous structure, and prevalence of active O-containing surface functional groups, as validated by BET, SEM, and FTIR analysis.

C.M. contributed to conceptualization, methodology, investigation, data curation, writing – original draft, review, and editing. M.B. contributed to methodology, investigation, data curation. A.A. and A.K.D. contributed to conceptualization and supervision.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Ahmad
R.
&
Hasan
I.
2016
Optimization of the adsorption of Pb (II) from aqueous solution onto PAB nanocomposite using response surface methodology
.
Environmental Nanotechnology, Monitoring & Management
6
(
December
),
116
129
.
https://doi.org/10.1016/j.enmm.2016.09.002
.
Al-Saydeh
S. A.
,
El-Naas
M. H.
&
Zaidi
S. J.
2017
Copper removal from industrial wastewater: A comprehensive review
.
Journal of Industrial and Engineering Chemistry
56
,
35
44
.
https://doi.org/10.1016/j.jiec.2017.07.026
.
Anastopoulos
I.
,
Pashalidis
I.
,
Hosseini-Bandegharaei
A.
,
Giannakoudakis
D. A.
,
Robalds
A.
,
Usman
M.
,
Escudero
L. B.
,
Zhou, Y., Colmenares, Y. C., Núñez-Delgado, A. & Lima, E. C.
2019
Agricultural biomass/waste as adsorbents for toxic metal decontamination of aqueous solutions
.
Journal of Molecular Liquids
295
(
December
),
111684
.
https://doi.org/10.1016/j.molliq.2019.111684
.
Ayora
C.
,
Macías
F.
,
Torres
E.
,
Lozano
A.
,
Carrero
S.
,
Nieto
J.-M.
,
Pérez-López
R.
,
Fernández-Martínez
A.
&
Castillo-Michel
H.
2016
Recovery of rare earth elements and yttrium from passive-remediation systems of acid mine drainage
.
Environmental Science & Technology
50
(
15
),
8255
8262
.
https://doi.org/10.1021/acs.est.6b02084
.
Bashir
S.
,
Zhu
J.
,
Fu
Q.
&
Hu
H.
2018
Comparing the adsorption mechanism of Cd by rice straw pristine and KOH-modified biochar
.
Environmental Science and Pollution Research
25
(
12
),
11875
11883
.
https://doi.org/10.1007/s11356-018-1292-z
.
Bashir
M.
,
Mohan
C.
,
Tyagi
S.
&
Annachhatre
A.
2022
Copper removal from aqueous solution using chemical precipitation and adsorption by himalayan pine forest residue as biochar
.
Water Science and Technology
86
(
3
),
530
554
.
https://doi.org/10.2166/wst.2022.222
.
Bhattacharyya
S. C.
2016
Effect of H2O2 concentrations on copper removal using the modified hydrothermal biochar
.
Bioresource Technology
207
(
May
),
262
267
.
https://doi.org/10.1016/j.biortech.2016.02.032
.
Chang
C.-F.
,
Chang
C.-Y.
&
Tsai
W.-T.
2000
Effects of burn-off and activation temperature on preparation of activated carbon from corn Cob agrowaste by CO2 and steam
.
Journal of Colloid and Interface Science
232
(
1
),
45
49
.
https://doi.org/10.1006/jcis.2000.7171
.
Chen
X.
,
Chen
G.
,
Chen
L.
,
Chen
Y.
,
Lehmann
J.
,
McBride
M. B.
&
Hay
A. G.
2011
Adsorption of copper and zinc by biochars produced from pyrolysis of hardwood and corn straw in aqueous solution
.
Bioresource Technology
102
(
19
),
8877
8884
.
https://doi.org/10.1016/j.biortech.2011.06.078
.
Choudhary
V.
,
Patel
M.
,
Pittman
C. U.
&
Mohan
D.
2020
Batch and continuous fixed-bed lead removal using himalayan pine needle biochar: Isotherm and kinetic studies
.
ACS Omega
5
(
27
),
16366
16378
.
https://doi.org/10.1021/acsomega.0c00216
.
Crane
R. A.
&
Sapsford
D. J.
2018
Towards ‘Precision mining’ of wastewater: Selective recovery of Cu from acid mine drainage onto diatomite supported nanoscale zerovalent iron particles
.
Chemosphere
202
(
July
),
339
348
.
https://doi.org/10.1016/j.chemosphere.2018.03.042
.
Deng
Y.
,
Zhang
T.
&
Wang
Q.
2017
Biochar adsorption treatment for typical pollutants removal in livestock wastewater: A review
. In:
Engineering Applications of Biochar
.
InTech
.
https://doi.org/10.5772/intechopen.68253
.
Duan
S.
,
Huang
K.
&
Tao
M.
2022
Equilibrium and kinetic adsorption of hydrogen fluoride onto zeolite 3A
.
Adsorption Science & Technology
2022
(
August
),
1
18
.
Edited by Stefano Salvestrini. https://doi.org/10.1155/2022/5790408
.
EBC 2022 European Biochar Certificate - Guidelines for a Sustainable Production of Biochar. Version 10.1 from 10th Jan 2022. European Biochar Foundation (EBC), Arbaz, Switzerland.
http://european-biochar.org.
Edraki
M.
,
Golding
S. D.
,
Baublys
K. A.
&
Lawrence
M. G.
2005
Hydrochemistry, mineralogy and sulfur isotope geochemistry of acid mine drainage at the Mt. Morgan mine environment, Queensland, Australia
.
Applied Geochemistry
20
(
4
),
789
805
.
https://doi.org/10.1016/j.apgeochem.2004.11.004
.
Fakhri
A.
2017
Adsorption characteristics of graphene oxide as a solid adsorbent for aniline removal from aqueous solutions: Kinetics, thermodynamics and mechanism studies
.
Journal of Saudi Chemical Society
21
(
January
),
S52
S57
.
https://doi.org/10.1016/j.jscs.2013.10.002
.
Fan
Q.
,
Sun
J.
,
Chu
L.
,
Cui
L.
,
Quan
G.
,
Yan
J.
,
Hussain
Q.
&
Iqbal
M.
2018
Effects of chemical oxidation on surface oxygen-containing functional groups and adsorption behavior of biochar
.
Chemosphere
207
(
September
),
33
40
.
https://doi.org/10.1016/j.chemosphere.2018.05.044
.
Gebrewold
B. D.
,
Kijjanapanich
P.
,
Rene
E. R.
,
Lens
P. N. L.
&
Annachhatre
A. P.
2019
Fluoride removal from groundwater using chemically modified rice husk and corn Cob activated carbon
.
Environmental Technology
40
(
22
),
2913
2927
.
https://doi.org/10.1080/09593330.2018.1459871
.
Huang
S.-H.
&
Chen
D.-H.
2009
Rapid removal of heavy metal cations and anions from aqueous solutions by an amino-Functionalized magnetic nano-adsorbent
.
Journal of Hazardous Materials
163
(
1
),
174
179
.
https://doi.org/10.1016/j.jhazmat.2008.06.075
.
Jiwan
S.
&
Ajay
K.
2011
Effects of heavy metals on soil, plants, human health and aquatic life
.
International Journal of Research in Chemistry and Environment
.
Available from: www.ijrce.org.
Kalavathy
M. ,. H.
,
Regupathi
I.
,
Pillai
M. G.
&
Miranda
L. R.
2009
Modelling, analysis and optimization of adsorption parameters for H3PO4 activated rubber wood sawdust using Response Surface Methodology (RSM)
.
Colloids and Surfaces B: Biointerfaces
70
(
1
),
35
45
.
https://doi.org/10.1016/j.colsurfb.2008.12.007
.
Karthikeyan
T.
,
Rajgopal
S.
&
Miranda
L.
2005
Chromium(VI) adsorption from aqueous solution by sawdust activated carbon
.
Journal of Hazardous Materials
124
(
1–3
),
192
199
.
https://doi.org/10.1016/j.jhazmat.2005.05.003
.
Kefeni
K. K.
,
Msagati
T. A. M.
&
Mamba
B. B.
2017
Acid mine drainage: Prevention, treatment options, and resource recovery: A review
.
Journal of Cleaner Production
151
(
May
),
475
493
.
https://doi.org/10.1016/j.jclepro.2017.03.082
.
Kwak
J.-H.
,
Islam
M. S.
,
Wang
S.
,
Messele
S. A.
,
Anne Naeth
M.
,
El-Din
M. G.
&
Chang
S. X.
2019
Biochar properties and lead(II) adsorption capacity depend on feedstock type, pyrolysis temperature, and steam activation
.
Chemosphere
231
(
September
),
393
404
.
https://doi.org/10.1016/j.chemosphere.2019.05.128
.
Lan
Y.
,
Peng
P.
&
Shi
B.
2011
Optimization of preparation conditions for PDMS-Silica composite pervaporation membranes using response surface methodology
.
Separation Science and Technology
46
(
14
),
2211
2222
.
https://doi.org/10.1080/01496395.2011.594476
.
Langmuir
I.
1918
The adsorption of gases on plane surfaces of glass, mica and platinum
.
Journal of the American Chemical Society
40
(
9
),
1361
1403
.
https://doi.org/10.1021/ja02242a004
.
Larsson
M.
,
Nosrati
A.
,
Kaur
S.
,
Wagner
J.
,
Baus
U.
&
Nydén
M.
2018
Copper removal from acid mine drainage-polluted water using glutaraldehyde-polyethyleneimine modified diatomaceous earth particles
.
Heliyon
4
(
2
),
e00520
.
https://doi.org/10.1016/j.heliyon.2018.e00520
.
Lian
F.
,
Sun
B.
,
Chen
X.
,
Zhu
L.
,
Liu
Z.
&
Xing
B.
2015
Effect of Humic Acid (HA) on sulfonamide sorption by biochars
.
Environmental Pollution
204
(
September
),
306
312
.
https://doi.org/10.1016/j.envpol.2015.05.030
.
Luef
E.
,
Prey
T.
&
Kubicek
C. P.
1991
Biosorption of zinc by fungal mycelial wastes
.
Applied Microbiology and Biotechnology
34
(
5
),
688
692
.
https://doi.org/10.1007/BF00167924
.
Masindi
V.
,
Chatzisymeon
E.
,
Kortidis
I.
&
Foteinis
S.
2018
Assessing the sustainability of Acid Mine Drainage (AMD) treatment in South Africa
.
Science of The Total Environment
635
(
September
),
793
802
.
https://doi.org/10.1016/j.scitotenv.2018.04.108
.
Nie
T.
,
Hao
P.
,
Zhao
Z.
,
Zhou
W.
&
Zhu
L.
2019
Effect of oxidation-induced aging on the adsorption and co-adsorption of tetracycline and Cu2+ onto biochar
.
Science of The Total Environment
673
(
July
),
522
532
.
https://doi.org/10.1016/j.scitotenv.2019.04.089
.
Ofudje
E. A.
,
Akiode
O. K.
,
Oladipo
G. O.
,
Adedapo
A. E.
,
Adebayo
L. O.
&
Awotula
A. O.
2015
Application of raw and alkaline-modified coconut shaft as a biosorbent for Pb2+ removal
.
BioResources
10
(
2
).
https://doi.org/10.15376/biores.10.2.3462-3480
.
Ortega
P. F. R.
,
Trigueiro
J. P. C.
,
Santos
M. R.
,
Denadai
Â. M. L.
,
Oliveira
L. C. A.
,
Teixeira
A. P. C.
,
Silva
G. G.
&
Lavall
R. L.
2017
Thermodynamic study of methylene blue adsorption on carbon nanotubes using isothermal titration calorimetry: A simple and rigorous approach
.
Journal of Chemical & Engineering Data
62
(
2
),
729
737
.
https://doi.org/10.1021/acs.jced.6b00804
.
Osman, A. I., El-Monaem, E. M. A., Elgarahy, A. M., Aniagor, C. O., Hosny, M., Farghali, M., Rashad, E., Ejimofor, M. I., López-Maldonado, E. A., Ihara, I. & Yap, P. S.
2023
Methods to prepare biosorbents and magnetic sorbents for water treatment: A review
.
Environmental Chemistry Letters
21
(
4
),
2337
2398
.
https://doi.org/10.1007/s10311-023-01603-4
.
Peng
H.
,
Gao
P.
,
Chu
G.
,
Pan
B.
,
Peng
J.
&
Xing
B.
2017
Enhanced adsorption of Cu(II) and Cd(II) by phosphoric acid-modified biochars
.
Environmental Pollution
229
(
October
),
846
853
.
https://doi.org/10.1016/j.envpol.2017.07.004
.
Qiu
B.
,
Tao
X.
,
Wang
H.
,
Li
W.
,
Ding
X.
&
Chu
H.
2021
Biochar as a low-cost adsorbent for aqueous heavy metal removal: A review
.
Journal of Analytical and Applied Pyrolysis
155
(
May
),
105081
.
https://doi.org/10.1016/j.jaap.2021.105081
.
Ray
A.
2020
Characterization of biochars from various agricultural by-products using FTIR spectroscopy, SEM focused with image processing
.
International Journal of Agriculture Environment and Biotechnology
13
,
4
.
https://doi.org/10.30954/0974-1712.04.2020.6
.
Rodríguez-Galán
M.
,
Baena-Moreno
F. M.
,
Vázquez
S.
,
Arroyo-Torralvo
F.
,
Vilches
L. F.
&
Zhang
Z.
2019
Remediation of acid mine drainage
.
Environmental Chemistry Letters
17
(
4
),
1529
1538
.
https://doi.org/10.1007/s10311-019-00894-w
.
San Miguel
G.
,
Fowler
G. D.
&
Sollars
C. J.
2003
A study of the characteristics of activated carbons produced by steam and carbon dioxide activation of waste tyre rubber
.
Carbon
41
(
5
),
1009
1016
.
https://doi.org/10.1016/S0008-6223(02)00449-9
.
Shane
A.
,
Xu
X.
,
Siame
J.
,
Nguvulu
A.
,
Tena
T. M.
,
Lungu
M.
,
Chinyanta
S.
,
Kawala
J.
,
Bowa
V. M.
&
Chirambo
B.
2021
Removal of copper from Acid Mine Drainage (AMD) or Acid Rock Drainage (ARD)
.
Journal of Water Resource and Protection
13
(
07
),
435
454
.
https://doi.org/10.4236/jwarp.2021.137026
.
Shu
J.
,
Cheng
S.
,
Xia
H.
,
Zhang
L.
,
Peng
J.
,
Li
C.
&
Zhang
S.
2017
Copper loaded on activated carbon as an efficient adsorbent for removal of methylene blue
.
RSC Advances
7
(
24
),
14395
14405
.
https://doi.org/10.1039/C7RA00287D
.
Simate
G. S.
&
Ndlovu
S.
2014
Acid mine drainage: Challenges and opportunities
.
Journal of Environmental Chemical Engineering
2
(
3
),
1785
1803
.
https://doi.org/10.1016/j.jece.2014.07.021
.
Swiatkowski
A.
,
Pakula
M.
,
Biniak
S.
&
Walczyk
M.
2004
Influence of the surface chemistry of modified activated carbon on its electrochemical behaviour in the presence of lead(II) ions
.
Carbon
42
(
15
),
3057
3069
.
https://doi.org/10.1016/j.carbon.2004.06.043
.
Tan
Y.
,
Chen
M.
&
Hao
Y.
2012
High efficient removal of Pb (II) by amino-Functionalized Fe3O4 magnetic nano-particles
.
Chemical Engineering Journal
191
(
May
),
104
111
.
https://doi.org/10.1016/j.cej.2012.02.075
.
Trakal
L.
,
Šigut
R.
,
Šillerová
H.
,
Faturíková
D.
&
Komárek
M.
2014
Copper removal from aqueous solution using biochar: Effect of chemical activation
.
Arabian Journal of Chemistry
7
(
1
),
43
52
.
https://doi.org/10.1016/j.arabjc.2013.08.001
.
Wang
Y.
&
Liu
R.
2018
H2O2 treatment enhanced the heavy metals removal by manure biochar in aqueous solutions
.
Science of The Total Environment
628–629
(
July
),
1139
1148
.
https://doi.org/10.1016/j.scitotenv.2018.02.137
.
Wongrod
S.
,
Simon
S.
,
Guibaud
G.
,
Lens
P. N. L.
,
Pechaud
Y.
,
Huguenot
D.
&
van Hullebusch
E. D.
2018
Lead sorption by biochar produced from digestates: Consequences of chemical modification and washing
.
Journal of Environmental Management
219
,
277
284
.
https://doi.org/10.1016/j.jenvman.2018.04.108
.
Wongrod
S.
,
Simon
S.
,
van Hullebusch
E. D.
,
Lens
P. N. L.
&
Guibaud
G.
2019
Assessing arsenic redox state evolution in solution and solid phase during As(III) sorption onto chemically-treated sewage sludge digestate biochars
.
Bioresource Technology
.
https://doi.org/10.1016/j.biortech.2018.12.056
.
Zhang
Y.
,
Li
X.
,
Zheng
W.
,
Ye
Z.
&
Li
Y.
2013
Enhanced removal of Pb 2 + from water by adsorption onto phosphoric acid-modified PS-EDTA resin: Mechanism and kinetic study
.
Desalination and Water Treatment
51
(
37–39
),
7223
7235
.
https://doi.org/10.1080/19443994.2013.792523
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).