In this study, activated carbon (AC) was chemically activated using sodium hydroxide (NaOH), and polyethyleneimine (PEI) was grafted onto the AC using glutaraldehyde as a cross-linking agent. Then the modified AC was applied to treat water samples containing copper ions (Cu2+). Preparation of AC-NaOH@PEI. The grafted AC was characterized, demonstrating that the specific surface area of material decreased from 959.3 to 556.9 m2/g. The ζ-potential changed from −27.2 to 10.4 mV, and the presence of a distinct flocculation on the surface of the AC was observed via scanning electron microscopy. The results demonstrated that PEI was successfully grafted onto the surface of AC. Furthermore, the adsorption results indicated that the Cu2+ adsorption capacity of AC-NaOH@PEI was greatly enhanced with increasing PEI loading. The adsorption amount of Cu2+ by the grafted AC-NaOH@PEI-200 increased from 20.02 to 47.8 mg/g. In addition, the adsorption of Cu2+ by AC-NaOH@PEI was a pH dependent process. At a pH of 6, the maximum removal rate reached 93%. The adsorption process is better described by the Langmuir and quasi-second order adsorption models, signifying that the adsorption of Cu2+ on AC@PEI consists of monolayer adsorption and chemisorption. After four adsorption-desorption cycles, AC@PEI exhibited high adsorption capacity for Cu2+, indicating that it has good regeneration ability. It is a promising adsorbent material.

  • Activated carbon modified with glutaraldehyde cross-linked polyethyleneimine was successfully prepared.

  • The adsorption of copper ions by activated carbon increased from 20.04 mg/g to 47.8 mg/g.

  • After four times of elution, the removal rate of copper ions could reach 84%.

  • The adsorption process uses the chelation of copper ions by the amino group in the polyethyleneimine grafted onto the surface of the activated carbon.

Graphical Abstract

Graphical Abstract
Graphical Abstract

With rapid industrial development, many pollutants containing heavy metals have been introduced into natural water bodies in recent years. Heavy metal ions cannot be biodegraded and can exist for a long time and enter the human living environment through the food chain (Abdolali et al. 2015; Yagmur Goren et al. 2022). When the level of copper ions (Cu2+) exceeds the normal human tolerance range, it affects the internal organs of the human body, especially the liver and gallbladder, and may lead to liver cirrhosis, ascites, and other severe diseases (da Silva Alves et al. 2021). Methods for removing Cu2+ include high-efficiency binders, chemical precipitation, ion exchange, and adsorption. Adsorption is a widely used method because it is economical, convenient, and easy to apply (Wang et al. 2021).

Activated carbon (AC), magnetic carbon nanotubes and graphene, and resins are often used as adsorbents to adsorb heavy metals. AC is the most widely used adsorbent in wastewater treatment, possessing the advantage of good stability, large specific surface area, and high adsorption performance (Fang et al. 2018; Shahrashoub & Bakhtiari 2021). However, the technology of removing Cu2+ through only by physical adsorption using AC has inherent limitations, such as limited absorption capacity, poor reusability, high ash content, uneven distribution of micropores, and an insufficient number of surface functional groups to selectively adsorb heavy metal elements (Fiorati et al. 2020; He et al. 2021). Therefore, it is necessary to modify its surface structure to improve the ability of AC to adsorb Cu2+. Adsorbents containing hydrazine, amine, and thioamide groups can remove metal ions from aqueous solutions by chelation. In particular, the amine group is one of the most important functional groups for the chelation of Cu2+ (Hernández-Morales et al. 2012).

Polyethyleneimine (PEI) is an organic macromolecule with a high cationic charge density that exhibits a high selective adsorption capacity through chelation and electrostatic interaction. PEI is suitable for treating Cu2+ due to the presence of a large number of amino groups and its good reactivity and ease of functionalization (Chen et al. 2018). However, PEI is difficult to recycle and can cause secondary pollution of water bodies. Furthermore, it cannot be applied alone to remove water pollution. Liu & Huang (2011) used eggshell as raw material and glutaraldehyde (GA) as the cross-linking agent, and PEI functionalized adsorbent was prepared by the cross-linking of functional groups, such as amide and aldehyde groups. As a result, the adsorption amount of chromium ion reached 160 mg/g. Chen et al. studied a magnetic gel material composed of GA polyimide PEI-modified corn cob and applied it to the adsorption of heavy metal ions in an aqueous solution. They exhibited effective Cu2+ and lead ion adsorption at 303 K, with maximum adsorption of 459.4 and 290 mg/g, respectively (Chen et al. 2022). Deng & Ting (2005) developed a surface-modified biomass adsorbent by modifying with PEI. The adsorption capacity of the modified material to Cu2+, lead ion and nickel ion were significantly higher. The adsorption capacity of Cu2+, lead ion and nickel ion reached 92, 204, 55 mg/g.

In this study, PEI was firmly attached to the surface of AC applying GA as a cross-linking agent. The PEI-modified AC composite adsorbent, AC@PEI, was prepared for the adsorption of Cu2+ in wastewater. The optimal adsorption conditions, such as pH, initial concentration, and temperature during adsorption, were studied through experiments. Furthermore, the thermodynamics, kinetics, and adsorption mechanism of the Cu2+ adsorption process of the modified adsorption material were studied.

Materials and reagents

NaOH, PEI, HCl CuSO4·5H2O, and ethylenediaminetetraacetic acid disodium salt dihydrate (EDTA) were provided by Aladdin Industries Co. (Shanghai). GA (50%) was procured from Damao Chemical Reagent Co. (Tianjin). Tris-buffered saline (TBS) buffer (0.02 mol/L) was purchased from Sempega Biotechnology Co. (Nanjing). All the chemicals employed in this study were of analytical grade and used without further purification. AC was purchased from Ningxia Burtons Activated Carbon Co. (Yinchuan).

Preparation of the absorbents

First, the AC (5 g) was washed repeatedly with deionized water to remove dust and impurities from its surface. After washing, the AC was placed in an oven for drying. Then, 3 g of the AC was placed in a NaOH solution with a concentration of 2 mol/L and a volume of 100 mL, and the reaction was carried out at a speed of 300 rpm for 3 h at 25 °C. The resulting material was washed to neutrality and dried in an oven at 105 °C, this material was called AC-NaOH. Next, 1 g of the AC-NaOH was placed in a round-bottom flask containing GA (10%) and stirred at 300 rpm for 2 h at 50 °C (pH adjusted to 8 with TBS buffer). After the stirring was completed, the resulting material was washed and dried. The obtained sample was called GA-AC. Finally, 50, 100, and 200 mg of PEI were weighed and placed in a solution of ethanol (5 mL), sonicated for 30 min, and then dissolved in 45 mL of deionized water. 1 g of GA-AC was added separately to different masses of PEI solution, and the resulting mixture was placed in a water bath at 50 °C and stirred at 300 rpm for 3 h. The AC-NaOH cross-linked with PEI for the three different amounts of PEI were referred to as AC-NaOH@PEI-50, AC-NaOH@PEI-100, and AC-NaOH@PEI-200. Figure 1 presents a schematic diagram of the preparation process and reaction (Zhu et al. 2015; Xie et al. 2022).
Figure 1

Experimental procedure for the preparation of PEI@AC.

Figure 1

Experimental procedure for the preparation of PEI@AC.

Close modal

Characterization of AC-NaOH@PEI

A Fourier-transform infrared (FTIR) spectrometer (Thermo Nicolet Avatar 380 USA) was used to analyze the surface functional groups and bonding, and the data were collected over a transmission range of 4,000–500 cm−1. A scanning electron microscope (SEM, JSM–7500F, Jeol, Japan) was used to analyze the morphological characteristics. The adsorption properties of the materials were analyzed using an X-ray photoelectron spectrometer (XPS, ESCALAB Xi + XPS, Thermo Fisher Scientific, USA). The surface elemental composition and valence were investigated. N2 adsorption-desorption isotherms (ASAP2460, Micromeritics, USA) were used to determine the adsorbents specific surface areas, pore volumes, and pore size distributions. The specific surface areas of the adsorbents were calculated using the Brunauer-Emmett-Teller (BET) equation, and the pore volumes and pore size distributions were calculated using the Barrett-Joyner-Halenda (BJH) equation. ζ-potential analysis using a Malvern Zeta potentiometer was employed to analyze the adsorbents and determine their surface ζ-potentials.

Adsorption experiments

Batch adsorption experiments

In a batch adsorption experiment, 0.1 g of each of the four materials (AC, AC-NaOH@PEI-50, AC-NaOH@PEI-100, and AC-NaOH@PEI-200) were weighed and added to 150 mL of a solution of CuSO4-5H2O at a concentration of 64 mg/L, and then placed in a mechanical shaker at 298.15 K, 150 rpm in the batch adsorption experiment with samples taken at 30–60 min intervals. In the isothermal adsorption experiments, 0.1 g of AC-NaOH@PEI-200 was weighed and mixed in 100 mL of CuSO4-5H2O solution with a concentration of 20–100 mg and then placed in a mechanical shaker at three different temperatures of 298.15, 308.15 and 318.15 K, respectively. After 24 h of adsorption. The pH of the initial solution was adjusted using 0.01 mol/L HCl in the pH experiments. Inductively coupled plasma mass spectrometry (ICP-MS) is used to detect Cu2+ in solution and calculate its adsorption capacity according to Equation (1), and the removal rate was calculated using the following Equation (2):
(1)
where qe is the adsorption capacity, C0 and Ce are the initial concentration of Cu2+ in the solution and the concentration of Cu2+ at residual concentration of Cu2+ in adsorption equilibrium, m is the mass of the adsorbent, V is the volume of the adsorption solution.
(2)
where η is the adsorption efficiency, C0 is the initial concentration of Cu2+ in the solution, Ce is the equilibrium adsorption concentration of Cu2+.

Characterization of AC-NaOH@PEI

FTIR analysis

Figure 2 presents the FTIR spectra of AC-NaOH@PEI-200 before and after adsorption of Cu2+, labeled AC-a and AC-b, respectively. FTIR spectrum shows that there is a wide and strong absorption peak in the range of 3,200–3,600 cm−1, which is due to the broadening of the plate tensile vibration peak of phenol hydroxyl or alcohol hydroxyl. The absorption peak at 2,917 cm−1 can be attributed to the asymmetric stretching and vibration of CH2 (Maneerung et al. 2016), whereas the absorption peaks at 1,631 cm−1 and 1,383 cm−1 can be attributed to the presence of C = O and C-OH stretching vibrations, respectively, associated with the carboxyl group. The spectrum of the AC was very similar to that of AC@PEI-a. However, the enhanced band at 3,447 cm−1 after PEI modification can be attributed to the overlap of -NH and -OH stretching vibrations. In the FTIR spectrum of AC@PEI-a, a new absorption peak can be observed at 1,534 cm−1, which is a typical deformation pattern of the -OH group combined with the -NH2 group on the surface of AC after NaOH treatment (Sambaza et al. 2017). The most apparent difference between AC-a and AC@PEI-b was the disappearance of the -NH2 absorption band at 1,471 cm−1 after the adsorption of Cu2+. Another difference between the FTIR spectra of AC@PEI-a and AC@PEI-b was the appearance of the hydroxyl vibrational peak at 1,402 cm−1 in the spectrum of AC@PEI-b after adsorption of Cu2+. These results suggest that GA was the bridge for PEI grafting to the surface of AC, whereas the nitrogen atom of the -NH group was the adsorption site of Cu2+ on AC@PEI (Deng & Ting 2005).
Figure 2

FTIR spectra of AC and AC@PEI, where (a) AC-NaOH@PEI-200 (b) is AC-NaOH@PEI-200 after adsorption.

Figure 2

FTIR spectra of AC and AC@PEI, where (a) AC-NaOH@PEI-200 (b) is AC-NaOH@PEI-200 after adsorption.

Close modal

Scanning electron microscopic (SEM) images of AC-NaOH@PEI

In Figure 3, the SEM images display AC (Figure 3(a) and 3(b)) and AC-NaOH (Figure 3(c) and 3(d)). AC had many impurities and collapsed pores. After NaOH treatment, a rougher surface could be observed and the number of surface pores increased significantly. AC-NaOH had pores blocked by impurities, the macropore structure collapsed, and new pores were formed. The number of pores on the surface increased significantly, and the volume of pores also increased. AC-NaOH@PEI-200 (Figure 3(e) and 3(f)) had a surface covered with lumpy and flocculent material, which indicated that GA successfully cross-linked PEI to the surface of AC, which was more conducive to the adsorption of Cu2+. Figure 3(g) and 3(h) display the morphology and EDS images of AC-NaOH@PEI-200 after adsorption of Cu2+ ions. Figure 3 indicates that the surface of AC-NaOH@PEI-200 was covered with Cu2+ after adsorption, demonstrating that AC@PEI-200 was prepared successfully (Jiang et al. 2019).
Figure 3

Scanning electron micrographs of AC (a, b), NaOH-AC (c, d), AC-NaOH@PEI-200 (e, f), and AC-NaOH@PEI-200 after adsorption (g, h).

Figure 3

Scanning electron micrographs of AC (a, b), NaOH-AC (c, d), AC-NaOH@PEI-200 (e, f), and AC-NaOH@PEI-200 after adsorption (g, h).

Close modal

Specific surface area of the adsorbent

The specific surface area, adsorption site density, porosity, surface characteristics, and pore structure of the adsorbents are the main factors that affect their adsorption capacity. Figure 4 displays the N2 adsorption-desorption isotherms and pore size distributions of AC, AC-NaOH, and AC@PEI. The specific surface area and average pore size calculated by the BET and BJH equations are presented in Table 1 (Tsubota et al. 2015). According to the IUPAC classification of adsorption isotherms, the N2 adsorption-desorption isotherms of AC, AC-NaOH, and AC@PEI are type-IV isotherms. In the low-pressure section (P/P0 = 0–0.4), a slight inflection point indicates the formation of monolayer dispersion. In contrast, in the intermediate region (P/P0 = 0.4), a slight slope suggests the formation of multilayer dispersion. In the high-pressure region (P/P0 = 0.4–1.0), there is a clear H4 hysteresis loop that closes at P/P0 = 0.4, indicating the presence of mesopores and micropores in the sample. The pore size distribution curves indicate that the pore sizes of AC, AC-NaOH, and AC@PEI with different PEI loadings were 2 nm with a relatively uniform pore size distribution, indicating that the prepared AC-NaOH@PEI adsorbents were mesoporous.
Table 1

Texture properties of AC and AC-NaOH@PEI adsorbent

MaterialSpecific surface area (m2/g)Pore volume (cm3/g)Pore size (nm)
AC 959.3 0.22 2.84 
AC-NaOH 1,170 0.26 2.85 
AC-NaOH@PEI-50 852.7 0.17 2.19 
AC-NaOH@PEI-100 675.1 0.15 2.83 
AC-NaOH@PEI-200 556.9 0.11 2.18 
MaterialSpecific surface area (m2/g)Pore volume (cm3/g)Pore size (nm)
AC 959.3 0.22 2.84 
AC-NaOH 1,170 0.26 2.85 
AC-NaOH@PEI-50 852.7 0.17 2.19 
AC-NaOH@PEI-100 675.1 0.15 2.83 
AC-NaOH@PEI-200 556.9 0.11 2.18 
Figure 4

N2 adsorption-desorption isotherms (a) and pore distribution (b) for AC, AC-NaOH and AC-NaOH@PEI-50, AC-NaOH@PEI-100, and AC-NaOH@PEI-200.

Figure 4

N2 adsorption-desorption isotherms (a) and pore distribution (b) for AC, AC-NaOH and AC-NaOH@PEI-50, AC-NaOH@PEI-100, and AC-NaOH@PEI-200.

Close modal

Table 1 demonstrates that for AC, the specific surface area was 959.3 m2/g, the pore volume was 0.22 cm3/g, and the pore size was 2.84 nm. For AC-NaOH, the specific surface area was 1,169.9 m2/g, the pore volume was 0.26 cm3/g, and the pore size was 2.85 nm, which were slightly higher than those of AC. Compared with that of AC@PEI, the specific surface area of AC@PEI decreased sharply with an increase in the amount of loaded PEI. Furthermore, AC@PEI displayed a significant decrease in its pore size and volume. AC-NaOH@PEI-200 of specific surface area and pore volume compared to AC were reduced to 556.9 m2/g and 0.11 cm3/g, respectively, due to GA successfully cross-linking PEI to occupy a particular space in the AC pores. The impregnation and binding of PEI on the AC surface partially blocked the pores, indicating that PEI was successfully combined with AC, which is consistent with the above FTIR results (Xie et al. 2019).

XPS study on the chemisorption of metal ions on AC-NaOH@PEI

To further investigate the adsorption mechanism of Cu2+, the AC@PEI-200 before and after the adsorption of Cu2+ were analyzed by XPS in Figure 5. The AC@PEI-200 curve (Figure 5(a)) of C1s consists of three fitted peaks with binding energies of 286.02, 285.44, 285.06 and 284.50 eV for C-N, C-O and C-C, respectively. The AC@PEI (Figure 5(c)) of O1s curve consists of a C = O fitted peak with a binding energy of 531.21 eV for the N1s spectrum. Figure 5(b) and 5(e) shows a clear variation. Before the adsorption of Cu2+, AC-NaOH@PEI-200 observed peaks at 399.44 eV and 401.17 eV, which -NH causes CN group causes the formation on the surface of AC@PEI-200 (Xie et al. 2019). After the adsorption of Cu2+, a new peak is observed at 402.07 eV for AC@PEI-200. The result shows that after AC-NaOH@PEI-200 adsorbs Cu2+, the nitrogen atoms in the adsorbent exist in an oxidized state. This is due to the formation of the -NH metal complex during the adsorption process. That is, the lone pair of electrons in the nitrogen atom contributes to the shared bond between Cu2+ and the nitrogen atom, so the nitrogen atom reduces the electron cloud density and the binding energy increases. This means that the -NH2 group may interact with Cu2+ during adsorption. In summary, the possible adsorption mechanism of AC-NaOH@PEI-200 for Cu2+ adsorption may involve the chemical interaction between the homogeneous surface functional groups of AC-NaOH@PEI-200 and Cu2+. During the adsorption process, -OH and -NH2 groups may interact with Cu2+ to form a coordination bond.
Figure 5

XPS spectra, (a-c) AC-NaOH@PEI-200, (c-d) AC-NaOH@PEI-200 after adsorption of Cu2+.

Figure 5

XPS spectra, (a-c) AC-NaOH@PEI-200, (c-d) AC-NaOH@PEI-200 after adsorption of Cu2+.

Close modal

ζ potential analysis

Table 2 presents the ζ-potential of the samples tested at room temperature and a pH of 6.0. The ζ-potential of the AC surface in aqueous solution was −27.2 mV, whereas after loading PEI, the ζ-potential of the AC@PEI surface in aqueous solution increased to 10.4 mV. The ζ-potential in aqueous solution increased to 10.4 mV, which is because the amine group of the PEI (-NH or -NH2) could readily adsorb cations (H+) and exhibited a positive charge on the surface, which led to a significant increase in the ζ-potential of the AC surface. This result demonstrates that PEI was successfully loaded on the AC surface (Chen et al. 2015). After the adsorption of Cu2+, the ζ-potential of the AC surface changed to 12.5 mV (Tian et al. 2015).

Table 2

ζ Potential of the PEI@AC adsorbent

MaterialZate potential (mV)
AC −27.2 
AC-NaOH@PEI-200 10.4 
AC-NaOH@PEI-200-adsorbed Cu2+ 12.5 
MaterialZate potential (mV)
AC −27.2 
AC-NaOH@PEI-200 10.4 
AC-NaOH@PEI-200-adsorbed Cu2+ 12.5 

Adsorption study

This study examined different loading, pH, adsorption time, temperature, and other experimental conditions. The testing procedures and results for different experimental conditions are described below.

Effects of PEI loading amount on adsorption performance

As illustrated in Figure 6(a), when PEI was loaded on AC, the adsorption capacity of Cu2+ increased with increasing PEI loading. The unmodified AC exhibited low Cu2+ adsorption of 20.04 mg/g after 4 h of adsorption, whereas AC-NaOH@PEI-50 and AC-NaOH@PEI-100 exhibited Cu2+ adsorption of 35.64 and 38.74 mg/g, respectively. When PEI was increased to 200 mg, the adsorption reached the optimal value of 45.11 mg/g and the adsorption equilibrium time was reduced to 3 h. These results were caused by the activation of NaOH, which increased the number of functional groups and improved the Cu2+ adsorption capacity of AC@PEI. As the highest adsorption was achieved with a PEI loading of 200 mg, the AC-NaOH@PEI-200 material was used in the subsequent experiment (Xie et al. 2019).
Figure 6

(a) Dynamic adsorption with different PEI loadings. (Adsorption conditions, T = 298.15 K, C0 = 64 mg/L, dosage = 0.1 g, V = 150 mL, rpm = 150), (b) Removal rates at different pH values. (Adsorption conditions, T = 298.15 K, C0 = 64 mg/L, AC-NaOH@PEI-200 = 0.1 g, time = 360 min, rpm = 150), (c) Static adsorption of AC-NaOH@PEI-200. (Adsorption conditions, AC-NaOH@PEI-200 = 0.1 g, time = 720 min, V = 100 mL, rpm = 150).

Figure 6

(a) Dynamic adsorption with different PEI loadings. (Adsorption conditions, T = 298.15 K, C0 = 64 mg/L, dosage = 0.1 g, V = 150 mL, rpm = 150), (b) Removal rates at different pH values. (Adsorption conditions, T = 298.15 K, C0 = 64 mg/L, AC-NaOH@PEI-200 = 0.1 g, time = 360 min, rpm = 150), (c) Static adsorption of AC-NaOH@PEI-200. (Adsorption conditions, AC-NaOH@PEI-200 = 0.1 g, time = 720 min, V = 100 mL, rpm = 150).

Close modal

Effects of pH on the adsorption

The metal ions in a solution and the chemical state of the reactive groups on the adsorbent (degree of protonation) are influenced by the pH of the solution. By adjusting the pH, the charge distribution on the adsorbent surface can be effectively manipulated, which affects the affinity of the adsorbent for the target metal ions and thus the adsorption efficiency. An alkaline pH produces Cu(OH)2 flocculent precipitate, which has a significant effect on the adsorption efficiency and leads to inaccurate adsorption results. Therefore, the pH range for this experiment was set to 2–6. Figure 6(b) illustrates the effect of the pH value on the removal rate of Cu2+. The Cu2+ adsorption behavior of the AC-NaOH@PEI-200 adsorbent was closely related to the pH value. The removal rate was 48% at pH = 2. The Cu2+ removal rate of AC@PEI increased significantly from 70 to 91% when the pH was increased from 3 to 5. The highest Cu2+ removal rate of AC-NaOH@PEI-200 was 93% when pH = 6 (Wang et al. 2015a, 2015b). The amine groups on the adsorbent surface were protonated at low pH by a large amount of H+ present in the solution, thus implying that complexation could not occur. In addition, electrostatic repulsion of Cu2+ occurred, which led to a decrease in the adsorption capacity. When the pH increased, the amount of amine present in the adsorbent increased the free radical content of the solution, competitive adsorption of protons and Cu2+ was weakened, electrostatic repulsion was reduced, and ligand group of Cu2+ on the surface of the adsorbent increased (Wang et al. 2015a, 2015b). By studying the effect of the initial solution pH, it was found that the optimal pH for Cu2+ adsorption for AC-NaOH@PEI-200 was 6 (Saleh et al. 2017).

Effect of the initial concentration and temperature on Cu2+ adsorption

The effect of the initial concentration of Cu2+ on the adsorption was investigated for the adsorption of Cu2+ by AC-NaOH@PEI-200 at three different temperatures of 298.15, 308.15, and 318.15 K, as illustrated in Figure 6(c). Different initial concentrations had a significant effect on the adsorption of Cu2+ by AC-NaOH@PEI-200. As the Cu2+ concentration increased from 20 to 100 mg/L, the equilibrium adsorption capacity qe of AC-NaOH@PEI-200 increased from 18.76 to 47.8 mg/g at 298.15 K. This can be explained by the fact that a higher initial concentration of Cu2+ produced a higher gradient concentration of Cu2+, which increased the mass transfer rate and led to greater uptake of Cu2+, thus increasing the Cu2+ adsorption capacity of AC-NaOH@PEI-200. In addition, the adsorption of qe decreased to 41.68 mg/g, when the temperature was increased to 318.15 K, indicating that the adsorption of Cu2+ on AC-NaOH@PEI-200 is an exothermic process.

Adsorption kinetics

The adsorption capacity of AC can be described using a quasi-first-order model, which can be expressed using the following Equation (3):
(3)
where K1 is the rate constant of the quasi-second order reaction, t is the adsorption time, qt is the adsorption capacity at time t, and qe is the adsorption capacity at adsorption equilibrium (Al-Anber & Matouq 2008).
A quasi-second-order model is usually used to study the kinetics of liquid-phase adsorption, which can be expressed using the following Equation (4):
(4)
where K2 is the rate constant of the proposed secondary reaction.
The intraparticle diffusion model is shown in Equation (5):
(5)
where Kid is the rate constant of the intraparticle diffusion (Liu et al. 2017).
Figure 7(a) and 7(b) present the kinetic fitting results for the adsorption of Cu2+ by AC-NaOH@PEI-200, whereas Table 3 presents the resulting model characteristic parameters. The results indicate that the correlation coefficient R2 of the quasi-second order kinetic equation was the most significant (R2 > 0.999), and the calculated value of the equilibrium adsorption capacity agreed well with the actual value, which indicates that the experimental adsorption data could be fitted well by a quasi-second-order kinetic equation and that the rate-limiting step was likely chemisorption. The Cu2+ adsorption was caused by the chemical adsorption between the active groups (i.e., amine groups) on the AC-NaOH@PEI-200 surface and the metal ions.
Table 3

Constants of AC-NaOH@PEI-200 adsorption kinetics

modelK1 (min/g)K2 (g/mg/min)Qe (mg/g)R2Kid1 mg/gmin1/2Kid2 mg/gmin1/2Kid3 mg/g min1/2
AC-NaOH@PEI-200 Quasi-first-order dynamics 0.0218 N/A 32.38 0.9693 N/A N/A N/A 
Quasi-second-order dynamics N/A 0.0013 47.85 0.9995 N/A N/A N/A 
Internal diffusion N/A N/A N/A 0.9710 7.0993 1.0560 0.0371 
    0.9145    
     0.9513    
modelK1 (min/g)K2 (g/mg/min)Qe (mg/g)R2Kid1 mg/gmin1/2Kid2 mg/gmin1/2Kid3 mg/g min1/2
AC-NaOH@PEI-200 Quasi-first-order dynamics 0.0218 N/A 32.38 0.9693 N/A N/A N/A 
Quasi-second-order dynamics N/A 0.0013 47.85 0.9995 N/A N/A N/A 
Internal diffusion N/A N/A N/A 0.9710 7.0993 1.0560 0.0371 
    0.9145    
     0.9513    
Figure 7

(a) Quasi primary kinetic model diagram (b) Quasi secondary kinetic model diagram (c) Intraparticle diffusion diagram.

Figure 7

(a) Quasi primary kinetic model diagram (b) Quasi secondary kinetic model diagram (c) Intraparticle diffusion diagram.

Close modal

Figure 7(c) presents the relevant equation as a curve. If the relationship between qt and t1/2 is linear, the adsorption is governed by an internal diffusion process. The intraparticle diffusion curve was composed of three or more phases. The initial linear part indicates transient or external surface adsorption, demonstrating that macropore diffusion achieved mass transfer in the early stages of adsorption. This suggests that the process was divided into three phases. The first stage involves the membrane diffusion of Cu2+ from the aqueous solution to the outer surface of the adsorbent. The second stage involves the intraparticle distribution of Cu2+ from the surface to the internal pores. As the value of Ki1 was greater than Ki2, the intraparticle distribution was a gradual process. The low concentration of Cu2+ and the strong repulsion between Cu2+ and AC-NaOH@PEI-200 led to the slow distribution of AC-NaOH@PEI-200, indicating that the model was a good fit for the adsorption process (Al-Anber & Matouq 2008; Simonin 2016).

Adsorption isotherms

To investigate the process of the adsorption of Cu2+ by the AC-NaOH@PEI-200, adsorption isotherm experiments were conducted at initial Cu2+ concentrations of 20–100 mg/L. Langmuir and Freundlich isotherm models were then used to model the adsorption behavior, the equations of which are shown in Equations (6) and (7), respectively.
(6)
(7)
where Ce is the concentration of Cu2+ adsorbed at equilibrium, qe is the amount of Cu2+ adsorbed at equilibrium, qm is the maximum amount of Cu2+ adsorbed at equilibrium, and KL is the Langmuir constant related to the affinity of the binding sites.

where n is a constant representing the adsorption strength, KF is a constant related to the adsorption capacity, and 1/n is the heterogeneity factor related to the adsorption strength and heterogeneity of the surface of the material (Jang et al. 2018).

Figure 8(a) presents two isotherms of Cu2+ adsorption on AC-NaOH@PEI-200. According to Equation (6), Ce/qe versus Ce is a straight line, and KL and qm can be calculated from the intercept and slope, respectively. According to Equation (7), log qe versus log Ce is also a straight line, and KF and n can be calculated from the intercept and slope, respectively. Table 4 presents the correlation coefficients of the Langmuir and Freundlich models. The R2 values of the Equations (6) and (7) were used for comparison. The R2 values of Langmuir adsorption Eq were larger than those of Freundlich adsorption, indicating that the Langmuir adsorption model was more suitable than the Freundlich model for equilibrium data of Cu2+ adsorption. The results indicate that the adsorption of Cu2+ by AC-NaOH@PEI-200 was a monolayer adsorption process with specific uniform adsorption sites on the surface of AC-NaOH@PEI-200. The maximum Cu2+ theoretically adsorbed on AC-NaOH@PEI-200 was 57.47 mg/g. In addition, it can be observed from Table 4 that Cu2+ was adsorbed on the AC@PEI adsorbent mainly through coordination complexation with reactive group types (i.e., carboxylic acid and amine and imine groups), as previously described. Thus, the adsorption of Cu2+ on AC-NaOH@PEI-200 was a monolayer and followed the Langmuir model. The -NH2 and -COOH functional groups in the adsorbent played an important role in the adsorption process by sharing electron pairs to coordinate with the Cu2+ in the solution. The KL and KF values gradually decreased with increasing temperature, indicating that the adsorption was exothermic, and low temperature facilitated the reaction (Gu et al. 2019).
Table 4

Isotherm model parameters under different temperatures

TemperatureLangmuir modle
Freundlich model
KKL (L/mg)qm (mg/g)KF (mg/g)n
298.15 0.35 49.75 0.9923 19.08 3.80 0.9619 
308.15 0.09 54.34 0.9731 10.40 2.62 0.9689 
318.15 0.05 57.47 0.9625 6.400 2.04 0.9527 
TemperatureLangmuir modle
Freundlich model
KKL (L/mg)qm (mg/g)KF (mg/g)n
298.15 0.35 49.75 0.9923 19.08 3.80 0.9619 
308.15 0.09 54.34 0.9731 10.40 2.62 0.9689 
318.15 0.05 57.47 0.9625 6.400 2.04 0.9527 
Figure 8

(a) Plot of Langmuir adsorption isotherm under different temperatures, (b) plot of Freundlich adsorption isotherm under different temperatures, thermodynamic fitting of Cu2+ adsorbed on AC-NaOH@PEI-200.

Figure 8

(a) Plot of Langmuir adsorption isotherm under different temperatures, (b) plot of Freundlich adsorption isotherm under different temperatures, thermodynamic fitting of Cu2+ adsorbed on AC-NaOH@PEI-200.

Close modal

Adsorption thermodynamics

The adsorption thermodynamic parameters for the entropy change (ΔS), enthalpy change (ΔH), and Gibbs free energy change (ΔG) can be expressed by the following Equations (8) and (9):
(8)
(9)
where R is the universal gas constant (8.314 J/mol/K), T is the absolute temperature in K, and ΔG °, ΔH, and ΔS are the changes in the Gibbs free energy, enthalpy, and entropy, respectively. Kc is the ratio of the AC-NaOH@PEI-200 concentration on the adsorbent at equilibrium (qe) to the remaining AC-NaOH@PEI-200 concentration in solution at equilibrium (Ce). ΔH and ΔS were calculated from the slope and intercept of the linear plot of ln Kc versus 1/T. Based on previous studies (Milonjic 2007; Zhu et al. 2011), the values of Kc were calculated by multiplication by 1000, as presented in Table 5.
Table 5

Regarding the value of Kc

Concentration (mg/L)298.15 K308.15 K318.15 K
20 15,129 4,934.7 2,913.9 
30 9,101.0 2,973.5 1,624.7 
40 4,063.3 1,989.5 1,492.2 
50 3,029.0 1,573.3 1,344.1 
60 2,401.4 1,467.1 1,260.7 
70 2,125.0 1,352.9 1,178.0 
80 1,484.5 1,250.4 1,081.7 
90 1,133.7 977.20 856.81 
100 915.70 802.12 714.67 
Concentration (mg/L)298.15 K308.15 K318.15 K
20 15,129 4,934.7 2,913.9 
30 9,101.0 2,973.5 1,624.7 
40 4,063.3 1,989.5 1,492.2 
50 3,029.0 1,573.3 1,344.1 
60 2,401.4 1,467.1 1,260.7 
70 2,125.0 1,352.9 1,178.0 
80 1,484.5 1,250.4 1,081.7 
90 1,133.7 977.20 856.81 
100 915.70 802.12 714.67 

Table 6 presents the thermodynamic parameters of the adsorption process, and Figure 8(b) presents the relationship between ln K and 1/T. Table 6 illustrates that ΔG < 0 and ΔH < 0, indicating that the adsorption of Cu2+ on AC-NaOH@PEI-200 was an exothermic process, similar to other findings reported in the literature. ΔS had a negative value. Before the adsorption of Cu2+ on AC-NaOH@PEI-200, Cu2+ moved freely in the solution, leading to a chaotic state. However, after the adsorption of Cu2+ on AC-NaOH@PEI-200, Cu2+ was in a stationary state, and the order of the adsorption process increased, leading to a decrease in ΔS values during adsorption, negative values of ΔS (Zhu et al. 2011). Table 6 illustrates that the value of ΔG changed from negative to positive with increasing temperature at Cu2+ concentrations of 90 and 100 mg/L, indicating that the adsorption of Cu2+ on AC-NaOH@PEI-200 was a spontaneous process at low temperatures and a nonspontaneous process at high temperatures (Tran et al. 2016; Ghosal & Gupta 2015, 2017).

Table 6

Thermodynamic parameters for Cu2+ adsorption on AC-NaOH@PEI-200

G(KJ/mol)
Concentration (mg/L)298.15 K308.15 K318.15 KH (KJ/mol)S (J/(mol·K))
20 −6,604 −3,899 −2,608 −68,170 −210 
30 −5,386 −2,656 −1,169 −71,910 −230 
40 −3,412 −1,681 −974.0 −41,570 −130 
50 −2,705 −1,096 −706.0 −34,080 −108 
60 −2,144 −950.0 −560.0 −27,020 −90.0 
70 −1,827 −731.0 −414.0 −24,110 −76.0 
80 −950.0 −536.0 −194.0 −12,890 −41.0 
90 −731.0 480.0 389.0 −19,120 −63.0 
100 −194.0 536.0 828.0 −17,460 −56.0 
G(KJ/mol)
Concentration (mg/L)298.15 K308.15 K318.15 KH (KJ/mol)S (J/(mol·K))
20 −6,604 −3,899 −2,608 −68,170 −210 
30 −5,386 −2,656 −1,169 −71,910 −230 
40 −3,412 −1,681 −974.0 −41,570 −130 
50 −2,705 −1,096 −706.0 −34,080 −108 
60 −2,144 −950.0 −560.0 −27,020 −90.0 
70 −1,827 −731.0 −414.0 −24,110 −76.0 
80 −950.0 −536.0 −194.0 −12,890 −41.0 
90 −731.0 480.0 389.0 −19,120 −63.0 
100 −194.0 536.0 828.0 −17,460 −56.0 

Desorption and regeneration experiments

The recoverability of the adsorbent affects the production costs and performance. Therefore, increasing the recoverability of the adsorbent is of great importance in practical applications (Abdolali et al. 2015). To study the recoverability of the AC-NaOH@PEI-200 adsorbent, 0.1 mol/L HCl and 0.1 mol/L EDTA solutions were used to study AC-NaOH@PEI-200. The results are presented in Figure 9(a). In this experiment, 0.1 g of adsorption-saturated AC-NaOH@PEI-200 was added to 0.1 mol/L HCl (100 mL) and 0.1 mol/L EDTA (100 mL) solutions. Subsequently, the two mixtures were eluted in a constant-temperature shaker at room temperature. Desorption was performed in a constant-temperature shaker at room temperature for 6 h. The adsorption-desorption experiment was repeated four times. Figure 9(a) presents the results of the experiment, the adsorption rate decreased from 94 to 81.07% after four cycles of elution of 0.1 mol/L HCl solution, and the adsorption efficiency decreased by 12.93%. In contrast, the adsorption rate decreased from 94 to 84.34% after four cycles of elution of 0.1 mol/L EDTA solution, and the adsorption efficiency decreased by 9.66%. The decrease in the removal rate of Cu2+ from the solution may have occurred due to an irreversible reaction that occurred between active sites of AC-NaOH@PEI-200 and Cu2+, which demonstrates that EDTA had a strong complexing ability and could desorb the metal ions of AC@PEI-200 without destroying the original structure of the adsorbent. This indicates that 0.1 mol/L EDTA solution was a suitable regenerant for AC-NaOH@PEI-200, as AC-NaOH@PEI-200 retained an adsorption efficiency of more than 84.34% over four cycles. These performance results reveal that AC-NaOH@PEI-200 exhibited not only a high Cu2+ removal rate but also good regeneration, indicating that this adsorbent has high potential for use in practical applications.
Figure 9

(a) Desorption and regeneration experiments by HCl and EDTA solution to AC-NaOH@PEI-200. (Adsorption conditions, AC-NaOH@PEI-200 = 0.1 g, time = 720 min,V = 100 mL, T = 298.15 k), (b) Competitive adsorption (Adsorption conditions, T = 298.15 K, C0 = 50 mg/L, AC@PEI-200 = 0.1 g, time = 360 min,V = 100 mL).

Figure 9

(a) Desorption and regeneration experiments by HCl and EDTA solution to AC-NaOH@PEI-200. (Adsorption conditions, AC-NaOH@PEI-200 = 0.1 g, time = 720 min,V = 100 mL, T = 298.15 k), (b) Competitive adsorption (Adsorption conditions, T = 298.15 K, C0 = 50 mg/L, AC@PEI-200 = 0.1 g, time = 360 min,V = 100 mL).

Close modal

Competitive adsorption

In the actual environment, wastewater contains more complex heavy metal ions. Therefore, it is necessary to investigate competitive adsorption. AC-NaOH@PEI-200 adsorbent (0.1 g) was added to an aqueous solution containing 50 mg/L of Cd2+, Li+, Cu2+ to form a competitive adsorption experiment of Cd2+ and Cu2+ or Li+. The results of competitive adsorption are shown in Figure 9(b). It is shows that the presence of Li+ has a significant inhibitory effect on the adsorption of Cu2+ on AC-NaOH@PEI-200, indicating that Cu2+ exhibits a strong competitive adsorption ability associated with Li+, which may be related to the size of the metal ion radius. The smaller the ionic radius, the easier it is to adsorb on AC-NaOH@PEI-200. The ionic radius of Li+ is 0.076 nm, and that of Cu2+ is 0.077 nm, so the adsorption capacity of AC-NaOH@PEI-200 is similar for Li+ and Cu2+. The ionic radius of Cu2+ is less than 0.95 for Cd2+, and the competition coefficient of Cd2+ is relatively tiny in the presence of Cu2+. Therefore, it is challenging to remove Cd2+ using AC-NaOH@PEI-200 adsorbent when Cu2+ is also present in the aqueous solution (Naghizadeh 2015; Deng et al. 2017).

Comparison of adsorption capacity

Adsorption capacity is a crucial index to evaluate the performance of materials. Some materials are listed in Table 7. Researcher studied that palm shell AC was modified by PEI as adsorbent to treat Cu2+ in waste water and the adsorption capacity reached 25.5 mg/g (Yin et al. 2007). Low-cost and effective AC for Cu2+ adsorption was prepared from walnut shell by CO2 activation in a fluidized bed, and the adsorption capacity reached 32.2 mg/g at 40 min. (Wu et al. 2018). Biochar was modified as a high efficient and selective absorbent for Cu2+ by nitration and reduction and the adsorption capacity reached 17.01 mg/g (Yang & Jiang 2014). In this work, the adsorption capacity of AC-NaOH@PEI-200 could reach 47.85 mg/g. AC-NaOH@PEI-200 shows the good adsorption capacity to Cu2+, indicating that it has great potential in the field of rapid wastewater treatment.

Table 7

Comparison of the adsorption of Cu2+ by different materials

Raw materialsModification methodAdsorption capacity (mg/g)Ref.
Charcoal PEI direct immersion 25.5 Yin et al. (2007)  
Walnut shell CO2 modification 32.2 Wu et al. (2018)  
Biological carbon Nitro reduction method 17.01 Yang & Jiang (2014
Commercial AC PEI cross-linking modification 47.8 This work 
Raw materialsModification methodAdsorption capacity (mg/g)Ref.
Charcoal PEI direct immersion 25.5 Yin et al. (2007)  
Walnut shell CO2 modification 32.2 Wu et al. (2018)  
Biological carbon Nitro reduction method 17.01 Yang & Jiang (2014
Commercial AC PEI cross-linking modification 47.8 This work 

In this study, GA was used to cross-link linear PEI, and the resulting of AC@PEI composites was used to adsorb and remove Cu2+ from wastewater. The specific surface area and pore volume of AC@PEI decreased significantly after surface loading of PEI. However, the Cu2+ adsorption capacity of AC-NaOH@PEI increased significantly. This indicates that the specific surface area of AC@PEI is not the decisive factor affecting the adsorption capacity. The surface ζ-potential of AC-NaOH@PEI-200 increased from −27.2 to 12.5 mV, and the surface became more positively charged. The pH value is an important factor affecting the adsorption of Cu2+. AC-NaOH@PEI-200 exhibited the most removal of Cu2+ in aqueous solutions when the pH was 6. The adsorption capacity of the untreated AC was only 20.02 mg/g, and the adsorption equilibrium required 4 h. The adsorption capacity of AC-NaOH@PEI-200 was increased to 47.8 mg/g and the adsorption equilibrium was shortened to 2 h. The adsorption isotherm data were in good agreement with the Langmuir adsorption equation, and the adsorption kinetics could be better described by a quasi-second order model, indicating that the adsorption of Cu2+ by AC-NaOH@PEI-200 involves monomolecular layer adsorption and that this adsorption is a chemical process. XPS analysis demonstrated that -NH2 and -NH were mainly involved in the adsorption of Cu2+. Combined with the cyclic regeneration results of AC-NaOH@PEI-200, it was found that the adsorption of Cu2+ on AC@PEI is a nonphysical process. i.e., a strong acid (0.1 mol/L HCl) and complexing solid agent (0.1 mol/L EDTA) are adequate for the regeneration of AC-NaOH@PEI-200, and the removal of Cu2+ adsorbed by AC@PEI retained 84% after four regeneration cycles. AC-NaOH@PEI-200 is thus a promising adsorption material that can be commercially mass-produced via a simple modification based on commercial AC.

Ningxia Hui Autonomous Region Key R&D Project 2022ZDYF0189 This work was financially supported by the Industrial Bioprocess Key Materials and Key Technology Innovation Team (No. KJT2019004).

The authors declare that they have no known competing financial interests or personal relationships that could have influenced the work reported in this paper.

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

The authors declare there is no conflict.

Abdolali
A.
,
Ngo
H. H.
,
Guo
W.
,
Zhou
J. L.
,
Du
B.
,
Wei
Q.
,
Wang
X. C.
&
Nguyen
P. D.
2015
Characterization of a multi-metal binding biosorbent: chemical modification and desorption studies
.
Bioresour. Technol.
193
,
477
487
.
da Silva Alves
D. C.
,
Healy
B.
,
Pinto
L. A. A.
,
Cadaval
T. R. S.
Jr.
&
Breslin
C. B.
2021
Recent developments in chitosan-based adsorbents for the removal of pollutants from aqueous environments
.
Molecules
26
,
594
.
Deng
J.
,
Liu
Y.
,
Liu
S.
,
Zeng
G.
,
Tan
X.
,
Huang
B.
,
Tang
X.
,
Wang
S.
,
Hua
Q.
&
Yan
Z.
2017
Competitive adsorption of Pb(II), Cd(II), and Cu(II) onto chitosan-pyromellitic dianhydride modified biochar
.
J. Colloid Interface Sci.
506
,
355
364
.
Fiorati
A.
,
Grassi
G.
,
Graziano
A.
,
Liberatori
G.
,
Pastori
N.
,
Melone
L.
,
Bonciani
L.
,
Pontorno
L.
,
Punta
C.
&
Corsi
I.
2020
Eco-design of nanostructured cellulose sponges for sea-water decontamination from heavy metal ions
.
J. Cleaner Prod.
246
,
10
.
Gu
S.-Y.
,
Hsieh
C.-T.
,
Gandomi
Y. A.
,
Yang
Z.-F.
,
Li
L.
,
Fu
C.-C.
&
Juang
R.-S.
2019
Functionalization of activated carbons with magnetic iron oxide nanoparticles for removal of copper ions from aqueous solution
.
J. Mol. Liq.
277
,
499
505
.
He
X.
,
Zhang
T.
,
Xue
Q.
,
Zhou
Y.
,
Wang
H.
,
Bolan
N. S.
,
Jiang
R.
&
Tsang
D. C. W.
2021
Enhanced adsorption of Cu(II) and Zn(II) from aqueous solution by polyethyleneimine modified straw hydrochar
.
Sci. Total Environ.
778
,
146116
.
Hernández-Morales
V.
,
Nava
R.
,
Acosta-Silva
Y. J.
,
Macías-Sánchez
S. A.
,
Pérez-Bueno
J. J.
&
Pawelec
B.
2012
Adsorption of lead (II) on SBA-15 mesoporous molecular sieve functionalized with -NH2 groups
.
Microporous Mesoporous Mater.
160
,
133
142
.
Liu
B.
&
Huang
Y.
2011
Polyethyleneimine modified eggshell membrane as a novel biosorbent for adsorption and detoxification of Cr(VI) from water
.
J. Mater. Chem.
21
,
9
10
.
Sambaza
S. S.
,
Masheane
M. L.
,
Malinga
S. P.
,
Nxumalo
E. N.
&
Mhlanga
S. D.
2017
Polyethyleneimine-carbon nanotube polymeric nanocomposite adsorbents for the removal of Cr6+ from water
.
Phys. Chem. Earth Parts A/B/C.
100
,
236
246
.
Tian
X.
,
Wang
W.
,
Wang
Y.
,
Komarneni
S.
&
Yang
C.
2015
Polyethylenimine functionalized halloysite nanotubes for efficient removal and fixation of Cr (VI)
.
Microporous Mesoporous Mater.
207
,
46
52
.
Tsubota
T.
,
Morita
M.
,
Kamimura
S.
&
Ohno
T.
2015
New approach for synthesis of activated carbon from bamboo
.
J. Porous Mater.
23
,
349
355
.
Wang
J.
,
Lu
X.
,
Ng
P. F.
,
Lee
K. I.
,
Fei
B.
,
Xin
J. H.
&
Wu
J. Y.
2015a
Polyethylenimine coated bacterial cellulose nanofiber membrane and application as adsorbent and catalyst
.
J. Colloid Interface Sci.
440
,
32
38
.
Wu
L.
,
Shang
Z.
,
Wang
H.
,
Wan
W.
,
Gao
X.
,
Li
Z.
&
Kobayashi
N.
2018
Production of activated carbon from walnut shell by CO2 activation in a fluidized bed reactor and its adsorption performance of copper ion
.
J. Mater. Cycles Waste Manage.
20
,
1676
1688
.
Xie
X.
,
Gao
H.
,
Luo
X.
,
Su
T.
,
Zhang
Y.
&
Qin
Z.
2019
Polyethyleneimine modified activated carbon for adsorption of Cd(II) in aqueous solution
.
J. Environ. Chem. Eng.
7
,
103183
.
Xie
X.
,
Gao
H.
,
Luo
X.
,
Zhang
Y.
,
Qin
Z.
&
Ji
H.
2022
Polyethyleneimine-modified magnetic starch microspheres for Cd(II) adsorption in aqueous solutions
.
Adv. Compos. Hybrid Mater.
5, 2772–2786.
Yagmur Goren
A.
,
Recepoglu
Y. K.
,
Karagunduz
A.
,
Khataee
A.
&
Yoon
Y.
2022
A review of boron removal from aqueous solution using carbon-based materials: an assessment of health risks
.
Chemosphere
293
,
133587
.
Zhu
H. Y.
,
Fu
Y. Q.
,
Jiang
R.
,
Jiang
J. H.
,
Xiao
L.
,
Zeng
G. M.
,
Zhao
S. L.
&
Wang
Y.
2011
Adsorption removal of Congo red onto magnetic cellulose/Fe3O4/activated carbon composite: equilibrium, kinetic and thermodynamic studies
.
Chem. Eng. J.
173
,
494
502
.
Zhu
D.
,
Li
X.
,
Liao
X.
&
Shi
B.
2015
Polyethyleneimine-grafted collagen fiber as a carrier for cell immobilization
.
J. Ind. Microbiol. Biotechnol.
42
,
189
196
.
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/).