The entry of heavy metals due to industrial activities into the environment is one of the major problems in this century. Nickel and chromium(VI) are the toxin elements that are used in various industries. In this study, lignocellulose nanofiber was purchased from the Nano-Novin Polymer Company (Gorgan, Iran) and used as an adsorbent for the removal of nickel and chromium(VI) ions from aqueous solutions in a batch system. The effects of pH, initial concentration, adsorbent dose, contact time, and temperature were investigated. The adsorption mechanism was determined by Langmuir, Freundlich, kinetic, and thermodynamic models. Investigation of equilibrium isotherms nickel and chromium(VI) showed that the isotherm fitted well with the Freundlich model. The pseudo-second-order model with the larger correlation coefficient had a greater fitness against experimental data in the kinetic studies. Thermodynamic parameters of both nickel and chromium such as Gibbs free energy, enthalpy, and entropy were calculated which indicated spontaneous, endothermic, and random processes, respectively. Lignocellulose nanofiber can be suggested as a good adsorbent that is highly capable of adsorbing nickel and chromium(VI) from aqueous solutions.

  • Lignocellulose nanofiber was purchased from the Nano-Novin Polymer Company.

  • Lignocellulose nanofiber as a good adsorbent is highly capable of adsorbing nickel and chromium from aqueous solutions.

  • The effects of pH, initial concentration, adsorbent dose, contact time, and temperature were investigated.

  • The adsorption mechanism was determined by Langmuir, Freundlich, kinetic, and thermodynamic models.

Nowadays, with increasing environmental problems and concerns, caused by heavy metal pollution, various industries such as metallurgical plants, plating, alloys, and battery production, technologies to treat polluted water should be developed rapidly (Rodiguez et al. 2018). One of the most important environmental concerns is the pollution caused by the toxicity of heavy metal ions, which accumulates in the food chain and has a high shelf life in nature (Xavier et al. 2018). The use of heavy metals in various industries is increasing which leads to environmental pollution (Maleki et al. 2019) Heavy metal ions are found naturally in volcanic activity, and weathering of rocks, as well as in many industries. The discharge of metal-polluted wastewater (e.g., brine) degrades water quality and thus water cannot be directly used for potable water (via desalination) and industrial applications (Panagopoulos 2022a, 2022b; Panagopoulos & Giannika 2022).

There are various methods for the purification of heavy metal ions in an aqueous solution, including chemical deposition, ion exchange, membrane filtration, precipitation, electrolytic method, reverse osmosis, solvent extraction, and adsorption (Manjuladevi et al. 2018; Anand et al. 2019; Kumar et al. 2019a, 2019b). Among all these methods, the adsorption method is preferred to other methods due to proper filtration, easy efficiency, and availability of various adsorbents (Zhang et al. 2018). The type of adsorbent and its method is essential in the adsorption method. In recent decades, the study of adsorbents has been expanded because they are highly effective, cheap, and economical (Kong et al. 2018).

The wide spread of nickel as a toxic heavy metal in the environment is harmful to human health. Ni(II) pollution in the aquatic environment is mainly caused by mines, melting, coal combustion, and plating (Yu et al. 2019). Once nickel enters the body, it penetrates all organs and accumulates in various tissues, causing damage to them, which may cause nerve, liver, kidney, and reproductive damage, and increase the risk of cancer (Shi et al. 2018). The high amount of nickel in the human body is dangerous to the skin (Zhou et al. 2018). Chromium is a natural element that is found mostly in nature in the form of trivalent and hexavalent chromium. Hexavalent chromium is more toxic than trivalent chromium (Arim et al. 2018). Chromium enters the human body through the skin, inhalation, or oral consumption and is transmitted by blood to the kidneys and liver. Hexavalent chromium is carcinogenic and damages DNA (Shi et al. 2018). Due to the high solubility of chromium in water, its accumulation is common in aqueous environments (Rizzo et al. 2019). Hexavalent chromium(VI) is a highly toxic heavy metal that is used in various industries such as leather, catalysts, fungicides, pigment production, ceramics, art, glass, photography, plating, etc. (Peng et al. 2019). Chromium causes problems such as bronchitis, cancer, wound formation, liver damage, weakened immune system, respiratory problems, and kidney damage (Gebru & Das 2018). The chromium(VI) concentration in drinking water used for human consumption may not be higher than 0.05 mg L−1, and the maximum allowable concentration in water and wastewater for the final discharge in the environment is 0.1 mg L−1 (Vilela et al. 2019).

Many inexpensive biopolymers such as chitosan, chitin, pectin, starch, cellulose, bacterial cellulose, lignin, and hemicellulose have been grafted with a variety of vinyl monomers for their use in separation technologies (Kumar & Sharma 2019) Cellulose is one of the most natural and plentiful renewable polymers (Kumar et al. 2019a, 2019b). Lingo cellulose is an ideal biological adsorbent because it is renewable, cheap, and has special structural properties. The main components of lignocelluloses are cellulose, hemicellulose, and lignin and it includes different kinds of functional groups such as hydroxyl, acetyl, phenolic, carboxyl, and methyl (An et al. 2018).

Removal of heavy metals through the application of various adsorbents has been of interest to many researchers because removal of this substance from wastewater before discharge into aquatic systems is necessary. Various technologies have been used in the past including chemical deposition, adsorption, ion exchange, membrane separation, solvent extraction, and biological processes to remove heavy metals from water and wastewater. Among these technologies, adsorption is the most appropriate method due to its high efficiency, low cost, convenient preparation, and operation. In the current study, the adsorption process was used to remove nickel and chromium(VI) ions. Adsorption is favored over other techniques due to ease of operation, low energy input, and removal of pollutants even at trace concentrations. The removal of nickel and chromium(VI) ions can be achieved by using various adsorbents like pomelo peel (Wu et al. 2017), Nano Kaolinite (Alasadi et al. 2019), Eichhornia crassipes and Lemna minor (Balasubramanian et al. 2019), microwave-functionalized cellulose (Qu et al. 2020), and coconut shell (Bal and Bhasarkar).

In this study, the application of lignocellulose nanofiber was investigated for the adsorption of nickel and hexavalent chromium from aqueous solutions in a batch system. The main objectives of this study were to investigate the effect of pH, initial concentration, contact time, temperature, and adsorbent dose on the adsorption of nickel and hexavalent chromium from aqueous solutions using lignocellulose nanofiber. For determining the mechanism of the adsorption in this study kinetic, thermodynamic, and adsorption isotherm models were investigated.

Lignocellulose nanofiber

Cellulose as a natural and renewable material has valuable physical properties which can be improved with detailed knowledge of the structure (Peter 2020). In addition, cellulose is the most abundant biomaterial and natural regenerating polymer (Ji et al. 2020). Lignocellulose nanofiber is composed of cellulose, hemicellulose, and lignin, and its raw materials include wood and other lignocellulose residues (e.g., wheat straw, rice straw, and sugarcane bagasse). This study used lignocellulose nanofiber for the adsorption of nickel and chromium(VI) ions from aqueous solutions. The advantages of this adsorbent over other adsorbents are specific surface area and ease of separation from the soluble phase (Zhang et al. 2020).

Cellulose nanofiber is typically produced using various mechanical methods such as homogenization, micro-fluidization, micro-grinding, refining, cryocrushing, or in combination with chemical and enzymatic pretreatments (Ferrer et al. 2012; Brodin et al. 2014). Many studies have been carried out on the use of cellulose nanofiber produced from virgin fiber as a strengthening agent for improving the physical and mechanical properties of paper, while the use of cellulose nanofiber isolated from bleached virgin fiber is not necessary or reasonable for many recycled/impure products (Yousefhashemi et al. 2019).

Lignocellulose nanofiber has been produced from cheap, recycled, old, corrugated container pulp by the use of an ultra-fine grinding technique. The diameter of the lignocellulose nanofiber is in the range of 10–80 nm, and the cellulose crystallinity index and crystallite size is 49% and 4 nm, respectively, during the process (Yousefhashemi et al. 2019).

Materials

This study was conducted on lignocellulose nanofiber as an adsorbent of heavy metal chromium(VI) and nickel ions from aqueous solutions on a laboratory scale in the Science and Research Branch of Azad University, Tehran, with two batch replications. Lignocellulose nanofiber was purchased from the Nano-Novin Polymer Company (Gorgan, Iran) with the physiochemical properties of lignocellulose nanofiber as shown in Table 1. The other chemicals were purchased from Merck Company (Germany).

Table 1

Properties of the Nano-Novin Polymer

NameLignocellulose nanofibers gel
Compounds Lignin, cellulose, hemicellulose 
Material state Gel 
Color Brown 
Production method Chemi-mechanical synthesis 
Nanofiber diameter Ave.: 55 nm 
NameLignocellulose nanofibers gel
Compounds Lignin, cellulose, hemicellulose 
Material state Gel 
Color Brown 
Production method Chemi-mechanical synthesis 
Nanofiber diameter Ave.: 55 nm 

Adsorption experiments

Batch adsorption studies were carried out to evaluate the efficiency of lignocellulose nanofiber adsorbent on the removal of nickel and chromium(VI) from the aqueous solutions. The stock solution of 1,000 mg L−1 was prepared by adding a certain amount of K2Cr2O7 and deionized water. The pH of each solution was adjusted using 0.1 M NaOH and 0.1 M HCl. Adsorption experiments were performed in the batch system. The pH of the solutions was measured using Crison pH Meter Basic 20. For agitation of the adsorbent and the heavy metal ions, the Heimolph Unimax 2010 shaker was used. For all treatments, the agitation speed was constant and equal to 120 rpm. After the reaction time, the lignocellulose nanofiber adsorbent was separated using a Kokusan H-108 centrifuge at 10 min and 4,000 rpm. Chromium(VI) and nickel ions concentration were measured using the Atomic Adsorption Device (AAS) model (Unicam-919). The effects of pH, adsorbent dose, contact time, initial concentration, and temperature were examined on the efficiency and adsorption capacity of nickel and chromium(VI) ions by the adsorbents. SPSS and Excel software were used for data analysis.

Analysis description and procedure

The experimental results were collected to identify the efficiency of adsorption criteria such as adsorption percentage, adsorption capacity, and breakthrough points at different operating conditions.

The adsorption percentage and adsorption capacity of the heavy metal chromium equilibrium were calculated using the following equations (Dehghani et al. 2017; Aslani et al. 2018):
(1)
(2)
In this formula, A is the percentage of adsorption; qe is the amount of adsorption capacity; C0 is the initial concentration of a metal ion in solution in milli grams per liter; Ce is the equilibrium concentration of metal ion in solution in milligrams per liter; m is the adsorbent mass in gram; and V is the volume of solution in liter.
The Langmuir equation is expressed by the following equation (Langmuir 1918):
(3)
(4)
In this equation, Ce (mg·L−1) is the equilibrium concentration of the metal ion in solution, qe is the equilibrium adsorption value (mg·g−1), qm is the maximum adsorption capacity, and b is the equilibrium constant of adsorption.
The Freundlich equation is expressed by the following equation (Veloso et al. 2020):
(5)
(6)
In this respect, qe is the amount absorbed in mg·g−1, Ce is the equilibrium concentration of the adsorbed ions in mg·L−1, and Kf is the Freundlich constant.

The study of biological adsorption at the temperature range produces thermodynamic constants such as the change of free energy of Gibbs (ΔG°), enthalpy change (ΔH°), and entropy change (ΔS°).

The free energy of Gibbs is expressed by the following equation (Ates & Basak 2019):
(7)
T refers to temperature (Kelvin), R is the ideal gas constant equal to 8.314 J·mol−1·K−1, and KC is the thermodynamic equilibrium constant obtained from the following formula:
(8)
Ca refers to mg of adsorbent, which absorbs per liter of solution, and Ce refers to soluble equilibrium concentration in milli grams per liter.
According to the laws of thermodynamics, the free energy of Gibbs at constant temperature depends on the enthalpy changes and the entropy changes expressed by Vant Hoff's Equation (9) (Kumar et al. 2018):
(9)
where ΔH° and ΔS° are obtained from the slope and intercept of the origin of the lnKC graph in terms of 1/T, respectively.
Pseudo-first-order kinetic is expressed by the following equation (Bao et al. 2020):
(10)
qt and qe are the amount of metal ions absorbed at time t (min) and at equilibrium time and k1 is the pseudo-first-order rate constant, respectively.
Pseudo-second-order kinetic is expressed by the following equation (Mahmoud et al. 2020):
(11)

Characterizations and properties of lignocellulose nanofiber

Fourier transform infrared spectroscopy (FTIR) and X-ray diffraction (XRD) were used to characterize the functional groups of the adsorbents and scanning electronic microscopy (SEM) was employed to observe the morphologies. Figure 1(a), the FTIR analysis was used to determine the superficial groups of lignocellulose nanofiber. As shown in Figure 3, in the FTIR spectrum of lignocellulose nanofibers, the wide peak (the left side) in the region of 2,800–3,600 cm−1 (3,359.81 cm–1) is related to the –OH (hydroxyl) and –NH functional groups in the FTIR spectrum by lignocellulose nanofibers. The peak in the region of 1,601.20 cm−1 is related to the functional group of C = O. The peak in the region of 1,028.29 cm−1 is related to the functional group of C–O. The next weak band in the region of 2,884.23 cm−1 after adsorption probably relates to the C–H functional groups. XRD analysis determines the crystalline structure of the adsorbent. As observed in Figure 1(b), the XRD curve of nano-crystalline peaks at 2θ is observed to be 16, 23, and 35.5, all indicating the crystalline structure of lignocellulose nanofiber. In order to study the diameter of lignocellulose nanofiber, transmission electron microscope (TEM) analysis was applied. According to Figure 1(c), the material used has less than 100 nm diameter (in the range of nanometers), and has a fibrous and network structure.
Figure 1

FTIR analysis (a), X-ray diffraction (b), and TEM (c) of lignocellulose nanofibers.

Figure 1

FTIR analysis (a), X-ray diffraction (b), and TEM (c) of lignocellulose nanofibers.

Close modal
Figure 2

The effect of pH on the adsorption percent-capacity of nickel and chromium(VI) ions by lignocellulose nanofibers (initial concentration = 10 mg/L, adsorbent dose = 0.1 g, contact time = 60 min, and T = 25 °C).

Figure 2

The effect of pH on the adsorption percent-capacity of nickel and chromium(VI) ions by lignocellulose nanofibers (initial concentration = 10 mg/L, adsorbent dose = 0.1 g, contact time = 60 min, and T = 25 °C).

Close modal
Figure 3

The effect of concentration on the adsorption percent-capacity of nickel (a) and hexavalent chromium (b) ions by lignocellulose nanofibers (adsorbent dose = 0.1 g, pH = 5, contact time = 60 min, and T = 25 °C).

Figure 3

The effect of concentration on the adsorption percent-capacity of nickel (a) and hexavalent chromium (b) ions by lignocellulose nanofibers (adsorbent dose = 0.1 g, pH = 5, contact time = 60 min, and T = 25 °C).

Close modal

Effect of pH

The solution pH has a considerable effect on the uptake of metal ions because it indicates the surface charge of the adsorbent and the degree of ionization and speciation of the adsorbate (Balasubramanian et al. 2019). In the adsorption process, pH has an important role in the removal of heavy metals (Barsbay et al. 2018; Esmaeili & Beni 2018). Adsorption experiments were carried out in the pH range of 2–7, keeping all other parameters constant (initial concentration = 10 mg/L, adsorbent dose = 0.1 g, contact time = 60 min, and T = 25 °C). For all treatments, the agitation speed was constant and equal to 120 rpm. When the pH rises above 7, the nickel and chromium hydroxides become insoluble, the adsorption capacity of nickel and chromium decreases, so the pH above 7 has not been investigated (Guo et al. 2018). The results of nickel and chromium(VI) are given in Figure 2(a) and 2(b), respectively. When the pH increases, the adsorption efficiency of nickel and chromium increases, so the adsorption percentage of nickel and chromium adsorption at pH = 5 are 91.2 and 83.4, respectively. Then at pH 6 and 7, the adsorption efficiency and adsorption capacity decrease. The results showed that the adsorption of nickel and chromium is pH-dependent, which affects the adsorbent surface properties and adsorption efficiency. Other researchers reported similar results for the adsorption of nickel and chromium by different adsorbents (An et al. 2019; Dehghani et al. 2019; Sethy & Sahoo 2019).

Statistical analysis of data by one-way ANOVA test indicated the directional effect of pH on the adsorption rate, which was significant (p < 0.05). Duncan's statistical test also showed that the optimum pH was 5. Duncan's test results also indicated significant differences in all cases, except the average pH of nickel at 5–7 and 2 and 3, the average pH of chromium at 4 and 6 and 4 and 7.

Effect of initial concentration of nickel and hexavalent chromium

The initial concentration is one of the important driving forces that affect the sorption process. The effect of changing the initial concentration of nickel and hexavalent chromium in the range of 5–50 mg/L while keeping the adsorbent dose (0.1 g), pH (6), contact time (60 min) and temperature (25 °C) constant are illustrated in Figure 5. For all treatments, the agitation speed was constant and equal to 120 rpm. According to Figure 3, the maximum adsorption percentages of nickel and chromium at concentration of 5 mg/L are 91.9 and 84.5%, respectively. However, at higher concentrations, the metal cations are far beyond the capacity of the receptors. As a result, competition for access to the contact area increases, and all connection sites are exposed to ions and are activated. The maximum adsorption capacity of nickel and chromium at concentration of 50 mg/L is 44.97 and 41.46, respectively. As the initial concentration increases, the adsorption capacity also increases (Aslan et al. 2018; Zhou et al. 2018). At low concentrations, nickel and chromium have more receptors to adsorb metal cations (Farokhi et al. 2018). As a result, nickel and chromium ions are simply able to bind to receptors on the surface, so the adsorption percentage increases. As can be seen in Figure 3, lignocellulose nanofibers have been investigated at different concentrations. The difference in the percentage of adsorption at the highest and lowest concentrations of nickel and chromium is small, which is 2 and 3% for nickel and chromium, respectively, which indicates good and satisfactory adsorption of adsorbent in different concentrations. Other researchers reported similar results for the adsorption of nickel and chromium by different adsorbents (Sivakumar et al. 2018; Mahmoud et al. 2020).

Statistical analysis of data by one-way ANOVA test indicated the directional effect of initial concentration on adsorption rate, which was significant (p < 0.05). Duncan test results of both nickel and chromium also indicated significant differences in all cases, except the average initial concentration of nickel at 5 and 30 and 10, 20, and 40 mg/g, and the average initial concentration of chromium at 30 and 40 mg/g.

Effect of adsorbent dose

The effect of adsorbent dose on nickel and chromium(VI) removal in the range of 0.1–0.6 g at fixed initial concentration nickel and chromium(VI) = 10 mg/L, pH = 5, contact time = 60 min, and T = 25 °C is shown in Figure 4. For all treatments, the agitation speed was constant and equal to 120 rpm. According to Figure 4, the maximum adsorption of nickel and chromium by lignocellulose nanofiber in the absorbent dose of 0.6 g is 96.65 and 92.7%, respectively. The adsorption capacity of the adsorbent dose is decreasing from 0.1 to 0.6 g, the maximum adsorption capacity of nickel and chromium in the adsorbent dose of 0.1 g is 9.12 and 8.34, respectively. The adsorption capacity declines with increasing adsorbent due to the increase in the number of unsaturated adsorption sites in the adsorbent (Sheikhi & Rezaei 2021). In general, the number of active adsorbent sites increases with the increase of adsorbent dose (Hadadian et al. 2018; Mousavi et al. 2019). As a result, the adsorption percentage increases. The maximum adsorption capacity of nickel and chromium at the adsorbent dose of 0.1 is 9.12 and 8.34, respectively, which indicates the high efficiency and cost-effectiveness of the adsorbent. As the adsorbent dose increases, the adsorption percentage of nickel and chromium ions increases because the number of adsorbable sites increases (Fan et al. 2019). Other researchers have found similar results for nickel and chromium uptake by different adsorbents (Hachoumi et al. 2019; Zhang et al. 2019).
Figure 4

The effect of adsorbent dosage on the adsorption percent-capacity of nickel (a) and hexavalent chromium (b) ions by lignocellulose nanofibers (initial concentration = 10 mg/L, pH = 5, contact time = 60 min, and T = 25 °C).

Figure 4

The effect of adsorbent dosage on the adsorption percent-capacity of nickel (a) and hexavalent chromium (b) ions by lignocellulose nanofibers (initial concentration = 10 mg/L, pH = 5, contact time = 60 min, and T = 25 °C).

Close modal
Figure 5

The effect of contact time on the adsorption percent-capacity of nickel (a) and hexavalent chromium (b) ions by lignocellulose nanofibers (initial concentration = 10 mg/L, pH = 5, adsorbent dose = 0.1 g, and T = 25 °C).

Figure 5

The effect of contact time on the adsorption percent-capacity of nickel (a) and hexavalent chromium (b) ions by lignocellulose nanofibers (initial concentration = 10 mg/L, pH = 5, adsorbent dose = 0.1 g, and T = 25 °C).

Close modal

Statistical analysis of data by one-way ANOVA test indicated the directional effect of adsorbent dose on adsorption rate, which was significant (p < 0.05). Duncan's statistical test also showed that the optimum adsorbent dose of nickel and chromium were 0.4 and 0.6 g, respectively. Duncan test results of both nickel and chromium also indicated significant differences in all cases, except the average adsorbent dose of nickel at 0.3, 0.4, and 0.5 and 0.2 and 0.3 g, and the average adsorbent dose of chromium at 0.1 and 0.2 g.

Effect of contact time

The effect of contact time was investigated within the range of 15–120 min, pH = 5, initial concentration of 5 mg·L−1 nickel and chromium, adsorbent dose of 0.1 g and temperature of 25 °C. As demonstrated in Figure 5, the adsorption increased with the test time. For all treatments, the agitation speed was constant and equal to 120 rpm. The highest adsorption rate of both nickel and chromium(VI) given as 96.4% was reached at 120 min. Figure 5 shows the effect of contact time on the adsorption percentage and adsorption capacity of nickel and chromium by lignocellulose nanofiber. As seen in Figure 5, the absorption rate increased from 15 to 120 min. Contact time must play an important role in the adsorption of heavy metals (Huang et al. 2018). Each substance must have a specific contact time to complete the adsorption process. Therefore, the adsorption property is highly dependent on the mixing time between the adsorbed substance and the adsorbent (Jiang et al. 2019). From the time of contact, it can be concluded that the amount of adsorption first increased and then reached equilibrium. Because of a larger number of active sites, the structure and area of the adsorbent are faster in the early times and allow the heavy metal to penetrate to a large extent (Bhowmik et al. 2018; Dim et al. 2021). Other researchers reported similar results for the adsorption of nickel and chromium by different adsorbents (Dim et al. 2021; Siuki et al. 2021).

In addition, statistical analysis of data by one-way ANOVA test indicated the directional effect of contact time on adsorption rate, which was significant (p < 0.05). Duncan's statistical test also showed that the optimum contact time of nickel and chromium were 60 and 90 min, respectively. Duncan test results of both nickel and chromium also indicated significant differences in all cases, except the average contact time of nickel at 60, 90, and 120 min, and the average contact time of chromium at 30 and 45 and 45 and 60 min.

Effect of temperature

Temperature plays a significant role in the chemical reaction between nickel and chromium(VI) with lignocellulose nanofiber. The adsorption experiments were conducted in the temperature range of 15–40 at constant nickel and chromium(VI) initial concentration (5 mg/L), pH (6), adsorbent dose (0.1 g/L) and contact time (60 min). For all treatments, the agitation speed was constant and equal to 120 rpm. The proper temperature can enhance the ion exchange and collision of atoms thereby improving the adsorption process (Shao et al. 2019). As can be seen in Figure 6, nickel adsorption by lignocellulose nanofiber increased from 54.75 to 97.15% with increasing temperature from 15 to 40 °C, and adsorption of Cr(VI) went up from 73.8 to 89.15% with the increase of temperature from 15 to 40 °C. According to Figure 6, adsorption went up with temperature because this adsorbent is not homogenous and the activation energies of adsorption sites vary. Therefore, at a low temperature, only the adsorption sites with low activation energy were occupied, while those with higher activation energy could be occupied only at higher temperatures (Kaewsichan & Tohdee 2019). The result obtained here is in agreement with the findings of other researchers who reported the adsorption of nickel and chromium by different adsorbents (Kundu et al. 2019; Mousavi et al. 2019).
Figure 6

The effect of temperature on the adsorption percent-capacity of nickel (a) and hexavalent chromium (b) ions by lignocellulose nanofibers (initial concentration = 10 mg/L, pH = 5, contact time = 60 min, and adsorbent dose = 0.1 g).

Figure 6

The effect of temperature on the adsorption percent-capacity of nickel (a) and hexavalent chromium (b) ions by lignocellulose nanofibers (initial concentration = 10 mg/L, pH = 5, contact time = 60 min, and adsorbent dose = 0.1 g).

Close modal

Statistical analysis of data by one-way ANOVA test indicated the directional effect of temperature on the adsorption rate, which was significant (p < 0.05). Duncan's statistical test also showed that the optimum temperature of nickel and chromium were 25 and 30 °C, respectively. Duncan test results of both nickel and chromium also indicated significant differences in all cases, except the average temperature of nickel at 25, 30, 35, and 40 °C, the average temperature of chromium at 20, 25, and 30, 30 and 35 °C, and 35 and 40 °C.

Adsorption isotherms

There are many models that can express the relationship between adsorption and residual concentration in solutions. Langmuir and Freundlich are more common and acceptable adsorption isotherms for the measurement of the adsorption of heavy metals in aqueous solutions (Hosseini et al. 2016). Therefore, in this study, Langmuir and Freundlich models have been applied to investigate and analyze experimental data and descriptions of the adsorption equilibrium between solid and liquid phases. The Langmuir model assumes uniform adsorption on the surface and is valid for monolayer adsorption with a homogeneous distribution of the adsorption sites and adsorption energies, while the Freundlich model shows multi-layer adsorption (Khelifi et al. 2018; Sun et al. 2019).

According to Figures 7 and 8 and Table 2 data, the Freundlich model with R2 = 0.9956 (nickel) and R2 = 0.9951 (chromium) compared to Langmuir with R2 = 0.5795 (nickel) and R2 = 0.2379 (chromium) can better describe the adsorption of nickel and chromium ions by lignocellulose nanofiber indicating that the data are consistent with the Freundlich model (Sohail et al. 2020). Maximum adsorption capacity (qmax) of nickel and chromium(VI) were equal to 270.271 and 344.827 mg·g−1, respectively, which indicates that the adsorbent has a relatively high adsorption capacity. Thus, the adsorbent can absorb heavy metals in high concentrations. The equilibrium constant of adsorption (b) of nickel and chromium(VI) was equal to 0.041 and 0.0147, respectively.
Table 2

Langmuir and Freundlich constants and coefficients in the adsorption of nickel and chromium(VI) using lignocellulose nanofibers

Heavy metalsFreundlich isotherm modelLangmuir isotherm model
Nickel Kf = 10.58 n = 1.07 R2 = 0.9956 b = 0.041 qmax = 270.271 R2 = 0.5795 
Chromium(VI) Kf = 5.17 n = 1.073 R2 = 0.9951 b = 0.0147 qmax = 344.827 R2 = 0.2379 
Heavy metalsFreundlich isotherm modelLangmuir isotherm model
Nickel Kf = 10.58 n = 1.07 R2 = 0.9956 b = 0.041 qmax = 270.271 R2 = 0.5795 
Chromium(VI) Kf = 5.17 n = 1.073 R2 = 0.9951 b = 0.0147 qmax = 344.827 R2 = 0.2379 
Figure 7

Langmuir adsorption isotherm curve for nickel (a) and hexavalent chromium (b) ions absorption using lignocellulose nanofibers.

Figure 7

Langmuir adsorption isotherm curve for nickel (a) and hexavalent chromium (b) ions absorption using lignocellulose nanofibers.

Close modal
Figure 8

Freundlich adsorption isotherm curve for nickel (a) and hexavalent chromium (b) ions absorption using lignocellulose nanofibers.

Figure 8

Freundlich adsorption isotherm curve for nickel (a) and hexavalent chromium (b) ions absorption using lignocellulose nanofibers.

Close modal

The calculated values of the Langmuir constant (KL) show a proper affinity of adsorption sites for nickel and chromium ions with lignocellulose nanofiber. Based on the calculated value of separation factor RL, its magnitude indicates the shape of the isotherm to be either favorable (0 < RL < 1), unfavorable (RL > 1), linear (RL = 1), or irreversible (RL = 0) (Al-Abbad et al. 2020).

As can be seen in Table 2, Kf is the Freundlich constant related to the bonding energy (L/g), and n is an empirical constant characterizing the heterogeneity of the process (g/L). According to the Freundlich model assumptions, if n < 1, the adsorption process is unfavorable, while if 1 < n > 10, the adsorption process is favorable. Results show that n in nickel and chromium were equal to 1.07 and 1.073, respectively, which indicated the favorable process for this adsorbent. Table 2 shows the constant coefficients and correlation coefficients of the Langmuir and Freundlich isotherms in chromium adsorption using lignocellulose nanofiber.

Since R2 is larger in Freundlich, the adsorption isotherm follows the Freundlich model. Freundlich isotherm is applicable to adsorption processes that occur on heterogonous surfaces. This isotherm gives an expression, which defines the surface heterogeneity and the exponential distribution of active sites and their energies.

Models of adsorption kinetics

The study of adsorption kinetics is essential to study the adsorption contact time (Mousavi et al. 2019). Adsorption kinetics is an important factor in the evaluation of adsorbents because it shows the amount of ions adsorbed by the adsorbent (Qureshi et al. 2017; Siddiqui et al. 2020). Adsorption kinetics models can be divided into two groups: reaction-based models and diffusion-based models. Reaction-based models include pseudo-first-order, pseudo-second-order, and Elovich models, and diffusion-based models include intraparticle diffusion, external-film diffusion, and internal-pore diffusion (Sarici-Ozdemir 2012). The models used in this study included reaction-based models of pseudo-first-order and pseudo-second-order.

The pseudo-first-order kinetic equation is based on adsorbent capacity and is used when adsorption occurs through the diffusion process inside a boundary layer (physical adsorption), while for the pseudo-second-order kinetic model, the model assumes that a metal ion is adsorbed to the available (two polar) sorption sites of the sorbent. The rate of pseudo-second-order sorption depends on the mass of the heavy metal ion on the sorbent surface and the number of heavy metal ions adsorbed at equilibrium (Aigbe & Osibote 2020).

Table 3 illustrates the pseudo-first-order kinetics and pseudo-second-order kinetics parameters in the adsorption of nickel and chromium using lignocellulose nanofiber. In order to evaluate the applicability of these kinetic models to fit the experimental data, K1 and K2 constants were determined experimentally from the slope and intercept of straight-line plots. As displayed in Figures 9 and 10, the results of the kinetic equation analysis showed that the adsorption data of nickel and chromium(VI) by lignocellulose nanofiber correspond to the pseudo-second-order model with R2 = 0.99. Thus, the adsorption data follow the pseudo-second-order kinetic model (Kumar et al. 2019a, 2019b).
Table 3

Pseudo-first-order kinetics and pseudo-second-order kinetics parameters in the adsorption of nickel and chromium(VI) using lignocellulose nanofibers

Pseudo-second-order kineticsPseudo-first-order kinetics
Nickel K = 0.038 R2 = 0.7681 K = 0.0873 R2 = 0.9941 
Chromium(VI) K = 0.0356 R2 = 0.6383 K = 0.0944 R2 = 0.9959 
Pseudo-second-order kineticsPseudo-first-order kinetics
Nickel K = 0.038 R2 = 0.7681 K = 0.0873 R2 = 0.9941 
Chromium(VI) K = 0.0356 R2 = 0.6383 K = 0.0944 R2 = 0.9959 
Figure 9

Pseudo-first-order adsorption kinetics adsorption in nickel (a) and hexavalent chromium (b) ions absorption using lignocellulose nanofibers.

Figure 9

Pseudo-first-order adsorption kinetics adsorption in nickel (a) and hexavalent chromium (b) ions absorption using lignocellulose nanofibers.

Close modal
Figure 10

Pseudo-second-order adsorption kinetics adsorption in nickel (a) and hexavalent chromium (b) ions absorption using lignocellulose nanofibers.

Figure 10

Pseudo-second-order adsorption kinetics adsorption in nickel (a) and hexavalent chromium (b) ions absorption using lignocellulose nanofibers.

Close modal

The analysis of the adsorption kinetics data according to the model Equations (10) and (11) resulted in the parameters listed in Table 3. The plot of ln(qeqt) vs. t should yield a straight line if the experimental data conform to this pseudo-first-order kinetic model. The plot of t/qt vs. t will give a straight line if the experimental data conform to this pseudo-second-order kinetic model, and the values of qe and K2 are obtained, respectively, from the slope and intercept of such a plot.

The results indicated that both the pseudo-first-order equation kinetic model and the pseudo-second-order kinetic model were followed, but it fits more with the second-order equation. In this model, the rate-limiting step is the surface adsorption that involves chemisorption, where the removal from a solution is because of physicochemical interactions between the two phases. The model is usually represented by its linear form.

Thermodynamics

Thermodynamic parameters were obtained by varying temperature conditions over the range of 10–40 °C by keeping other variables constant. The values of the thermodynamic parameters such as ΔG°, ΔH°, and ΔS°, describing nickel and chromium(VI) ions uptake by lignocellulose nanofibers, were calculated using the thermodynamic equations. The thermodynamic parameter indicates whether the adsorption reaction is spontaneous, random, exothermic, or endothermic.

Thermodynamic studies showed that the Gibbs free energy change (ΔG°) indicates whether a chemical reaction would take place spontaneously, negative ΔG° confirms spontaneous adsorption, and positive ΔG° indicates its prohibition (ΔG° < 0, ΔG° > 0), ΔH° (enthalpy change) is the amount of energy released (exothermic process) or consumed (endothermic process) by adsorption (ΔH° < 0, ΔH° > 0). A positive ΔS° (change in entropy) specifies that the adsorption comprises a dissociative mechanism, while a negative ΔS° corresponds to an associative mechanism (Kaewsichan & Tohdee 2019).

As Figure 11 and Tables 4 and 5 illustrate, the process of nickel and chromium(VI) ions removal by lignocellulose nanofibers was demonstrated to be possible in terms of stoichiometry, indicating that the adsorption process is spontaneous. As the temperature rises, ΔG° is reduced; as a result, the spontaneous reaction of the process was increased. Furthermore, the positive values of ΔH° (enthalpy) showed that the overall reaction process was endothermic, that is, the removal rate was declined by increasing the ambient temperature. The positive values of ΔS° (entropy) also indicated that the amount of irregularity was increased at the solid–liquid interface during the adsorption process. In fact, the positive value of ΔS° represents the affinity of adsorbent to adsorbate in the solution and some structural changes in adsorbent and adsorbate.
Table 4

Thermodynamic parameters of nickel adsorption using lignocellulose nanofibers

T (°C)ΔG° (kJ·mol−1)ΔH° (kJ·mol)ΔS (J·mol·k)
15 −454.942 96,134.782 +337.531 
20 −2774.606   
25 −5792.563   
30 −6514.501   
35 −7705.182   
40 −9180.85   
T (°C)ΔG° (kJ·mol−1)ΔH° (kJ·mol)ΔS (J·mol·k)
15 −454.942 96,134.782 +337.531 
20 −2774.606   
25 −5792.563   
30 −6514.501   
35 −7705.182   
40 −9180.85   
Table 5

Thermodynamic parameters of chromium(VI) adsorption using lignocellulose nanofibers

T (°C)ΔG° (kJ·mol−1)ΔH° (kJ·mol)ΔS (J·mol·k)
15 −2478.237 28,801.358 +109.586 
20 −3675.927   
25 −3998.801   
30 −4388.345   
35 −4957.538   
40 −5480.405   
T (°C)ΔG° (kJ·mol−1)ΔH° (kJ·mol)ΔS (J·mol·k)
15 −2478.237 28,801.358 +109.586 
20 −3675.927   
25 −3998.801   
30 −4388.345   
35 −4957.538   
40 −5480.405   
Figure 11

Thermodynamics in nickel (a) and hexavalent chromium (b) ions absorption using lignocellulose nanofibers.

Figure 11

Thermodynamics in nickel (a) and hexavalent chromium (b) ions absorption using lignocellulose nanofibers.

Close modal

The thermodynamics of nickel and chromium adsorption at various temperatures were studied according to Tables 4 and 5 where Gibbs free energy (ΔG°), enthalpy (ΔH°), and entropy (ΔS°) were calculated.

As Tables 4 and 5 show, thermodynamic parameters of nickel and chromium(VI) adsorption using lignocellulose nanofiber ΔG°, ΔH°, and ΔS° are indicators of the possible nature of adsorption, spontaneous, endothermic, and random, respectively. Other researchers reported similar results for the adsorption of nickel and chromium by different adsorbents (Adebayo et al. 2020; Mondal & Chakraborty 2020).

In this study, the application of lignocellulose nanofiber was investigated for the adsorption of hexavalent chromium and nickel from aqueous solutions in a batch system. This study evaluated the effect of pH, temperature, contact time, initial Cr and Ni concentration, and adsorbent dose parameters on the percentage of adsorption and absorption capacity. The results showed that a pH of 6, an adsorbent dose of 0.1 g at 25 °C and 60 min of contact time offered the optimum percentages of adsorption in this research. The lignocellulose nanofiber proved to be excellent for the removal of nickel and chromium metal ions from aqueous solutions not only at low concentrations but at high concentrations too. The results of the study of two-parameter isotherms (Langmuir and Freundlich) revealed that in general, the studied isotherms well predicted the equilibrium of the system. The adsorption data fitted well the Freundlich isotherm model, showing that the adsorption of nickel and chromium on the adsorbent was multi-layered. Kinetic studies also showed that the data corresponded to the pseudo-second-order model and the intraparticle. The calculation of thermodynamic parameters of chromium and nickel showed a spontaneous, endothermic, and random process. Maximum adsorption capacity (qmax) of nickel and chromium(VI) was equal to 270.271 and 344.827 mg·g−1, respectively, which indicates that the adsorbent has a relatively high adsorption capacity and cost-effectiveness of the adsorbent. Lignocellulose nanofiber, as the most abundant molecular biomass polymer in nature, has been characterized as possessing high biocompatibility, low toxicity, and high biodegradability properties. Thus, researchers are interested in such substances. According to the results of this study, lignocellulose nanofiber has a high potential for the adsorption of chromium and nickel from aqueous solutions.

The authors thank The Islamic Azad University Science and Research Branch of Iran for providing laboratory materials and facilities. I would like to acknowledge the help and support of a dear friend, Arash Azizi, who provided me with certain resources while writing this paper.

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

The authors declare there is no conflict.

Aigbe
U. O.
&
Osibote
O. A.
2020
A review of hexavalent chromium removal from aqueous solutions by sorption technique using nanomaterials
.
Journal of Environmental Chemical Engineering
8
(
6
),
104503
.
Al-Abbad
E.
,
Alakhras
F.
,
Anastopoulos
I.
,
Das
D.
,
AL-Arfaj
A.
,
Ouerfelli
N.
&
Hosseini-Bandegharaei
A.
2020
Chitosan-based materials for the removal of nickel ions from aqueous solutions
.
Russian Journal of Physical Chemistry A
94
,
748
755
.
Alasadi
A.
,
Khaili
F.
&
Awwad
A.
2019
Adsorption of Cu (II), Ni (II) and Zn (II) ions by nano kaolinite: thermodynamics and kinetics studies
.
Chemistry International
5
(
4
),
258
226
.
An
Y.
,
Zhang
X.
,
Wang
X.
,
Chen
Z.
&
Wu
X.
2018
Nano@ lignocellulose intercalated montmorillonite as adsorbent for effective Mn (II) removal from aqueous solution
.
Scientific Reports
8
(
1
),
1
11
.
Anand
S. R.
,
Aggarwal
R.
,
Saini
D.
,
Sonker
A. K.
,
Chauhan
N.
&
Sonkar
S. K.
2019
Removal of toxic chromium (VI) from the wastewater under the sunlight-illumination by functionalized carbon nano-rods
.
Solar Energy
193
,
774
781
.
Arim
A. L.
,
Guzzo
G.
,
Quina
M. J.
&
Gando-Ferreira
L. M.
2018
Single and binary sorption of Cr (III) and Ni (II) onto modified pine bark
.
Environmental Science and Pollution Research
25
(
28
),
28039
28049
.
Aslan
S.
,
Yildiz
S.
&
Ozturk
M.
2018
Biosorption of Cu2+ and Ni2+ ions from aqueous solutions using waste dried activated sludge biomass
.
Polish Journal of Chemical Technology
20
(
3
),
20
28
.
Ates
N.
&
Basak
A.
2019
Selective removal of aluminum, nickel and chromium ions by polymeric resins and natural zeolite from anodic plating wastewater
.
International Journal of Environmental Health Research
31
(
1
),
102
119
.
Balasubramanian
U. M.
,
Murugaiyan
S. V.
&
Marimuthu
T.
2019
Enhanced adsorption of Cr (VI), Ni (II) ions from aqueous solution using modified Eichhornia crassipes and Lemna minor
.
Environmental Science and Pollution Research
27
,
20648
20662
.
Barsbay
M.
,
Kavaklı
P. A.
,
Tilki
S.
,
Kavaklı
C.
&
Güven
O.
2018
Porous cellulosic adsorbent for the removal of Cd (II), Pb (II) and Cu (II) ions from aqueous media
.
Radiation Physics and Chemistry
142
,
70
76
.
Bhowmik
K. L.
,
Kanmani
M.
,
Deb
A.
,
Debnath
A.
,
Nath
R. K.
&
Saha
B.
2018
Mesoporous iron-manganese magnetic bimetal oxide for efficient removal of Cr (VI) from synthetic aqueous solution
.
Applied Mechanics and Materials
877
,
33
38
.
Brodin
F. W.
,
Gregersen
Ø. W.
&
Syverud
K.
2014
Cellulose nanofibrils: Challenges and possibilities as a paper additive or coating material–a review
.
Nordic Pulp & Paper Research Journal
29
(
1
),
156
166
.
Dehghani
M. H.
,
Niasar
Z. S.
,
Mehrnia
M. R.
,
Shayeghi
M.
,
Al-Ghouti
M. A.
,
Heibati
B.
,
McKay
G.
&
Yetilmezsoy
K.
2017
Optimizing the removal of organophosphorus pesticide malathion from water using multi-walled carbon nanotubes
.
Chemical Engineering Journal
310
,
22
32
.
Dim
P.
,
Mustapha
L.
,
Termtanun
M.
&
Okafor
J.
2021
Adsorption of chromium (VI) and iron (III) ions onto acid-modified kaolinite: isotherm, kinetics and thermodynamics studies
.
Arabian Journal of Chemistry
14
(
4
),
103064
.
Farokhi
M.
,
Parvareh
A.
&
Moraveji
M. K.
2018
Performance of ceria/iron oxide nano-composites based on chitosan as an effective adsorbent for removal of Cr (VI) and Co (II) ions from aqueous systems
.
Environmental Science and Pollution Research
25
(
27
),
27059
27073
.
Ferrer
A.
,
Quintana
E.
,
Filpponen
I.
,
Solala
I.
,
Vidal
T.
,
Rodríguez
A.
,
Laine
J.
&
Rojas
O. J.
2012
Effect of residual lignin and heteropolysaccharides in nanofibrillar cellulose and nanopaper from wood fibers
.
Cellulose
19
(
6
),
2179
2193
.
Guo
X.
,
Tang
S.
,
Song
Y.
&
Nan
J.
2018
Adsorptive removal of Ni2+ and Cd2+ from wastewater using a green longan hull adsorbent
.
Adsorption Science & Technology
36
(
1–2
),
762
773
.
Hachoumi
I.
,
Benkaddour
S.
,
El Ouahabi
I.
,
Slimani
R.
,
Cagnon
B.
,
El Haddad
M.
,
El Antri
S.
&
Lazar
S.
2019
Ensis siliqua shell for removal of Cu (II), Zn (II) and Ni (II) from aqueous solutions: Kinetics and isotherm model
.
Analytical Chemistry Letters
9
(
1
),
50
63
.
Hosseini
H.
,
Rezaei
H.
,
Shahbazi
A.
&
Maghsoodlu
A.
2016
Application of nano-lignocellulose for removal of nickel ions from aqueous solutions
.
Environmental Resources Research
4
(
2
),
213
229
.
Huang
H.
,
He
D.
,
Tang
Y.
,
Guo
Y.
,
Li
P.
,
Qv
W.
,
Deng
F.
&
Lu
F.
2018
Adsorption of hexavalent chromium from an aqueous phase by hydroxypropyl methylcellulose modified with diethylenetriamine
.
Journal of Chemical & Engineering Data
64
(
1
),
98
106
.
Jiang
D.
,
Yang
Y.
,
Huang
C.
,
Huang
M.
,
Chen
J.
,
Rao
T.
&
Ran
X.
2019
Removal of the heavy metal ion nickel (II) via an adsorption method using flower globular magnesium hydroxide
.
Journal of Hazardous Materials
373
,
131
140
.
Khelifi
O.
,
Nacef
M.
&
Affoune
A. M.
2018
Nickel (II) adsorption from aqueous solutions by physico-chemically modified sewage sludge
.
Iranian Journal of Chemistry and Chemical Engineering (IJCCE)
37
(
1
),
73
87
.
Kong
A.
,
Ji
Y.
,
Ma
H.
,
Song
Y.
,
He
B.
&
Li
J.
2018
A novel route for the removal of Cu (II) and Ni (II) ions via homogeneous adsorption by chitosan solution
.
Journal of Cleaner Production
192
,
801
808
.
Kumar
R.
,
Kim
S.-J.
,
Kim
K.-H.
,
Lee
S.-h.
,
Park
H.-S.
&
Jeon
B.-H.
2018
Removal of hazardous hexavalent chromium from aqueous phase using zirconium oxide-immobilized alginate beads
.
Applied Geochemistry
88
,
113
121
.
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
.
Mahmoud
M. E.
,
Amira
M. F.
,
Seleim
S. M.
&
Mohamed
A. K.
2020
Amino-decorated magnetic metal-organic framework as a potential novel platform for selective removal of chromium (Vl), cadmium (II) and lead (II)
.
Journal of Hazardous Materials
381
,
120979
.
Maleki
A.
,
Hajizadeh
Z.
,
Sharifi
V.
&
Emdadi
Z.
2019
A green, porous and eco-friendly magnetic geopolymer adsorbent for heavy metals removal from aqueous solutions
.
Journal of Cleaner Production
215
,
1233
1245
.
Panagopoulos
A.
2022b
Process simulation and analysis of high-pressure reverse osmosis (HPRO) in the treatment and utilization of desalination brine (saline wastewater)
.
International Journal of Energy Research
46
(
15
),
23083
23094
.
Peng
H.
,
Leng
Y.
&
Guo
J.
2019
Electrochemical removal of chromium (VI) from wastewater
.
Applied Sciences
9
(
6
),
1156
.
Peter
Z.
2020
Order in cellulosics: historical review of crystal structure research on cellulose
.
Carbohydrate Polymers
254
,
117417
.
Qu
J.
,
Tian
X.
,
Jiang
Z.
,
Cao
B.
,
Akindolie
M. S.
,
Hu
Q.
,
Feng
C.
,
Feng
Y.
,
Meng
X.
&
Zhang
Y.
2020
Multi-component adsorption of Pb (II), Cd (II) and Ni (II) onto microwave-functionalized cellulose: Kinetics, isotherms, thermodynamics, mechanisms and application for electroplating wastewater purification
.
Journal of Hazardous Materials
387
,
121718
.
Qureshi
M. I.
,
Patel
F.
,
Al-Baghli
N.
,
Abussaud
B.
,
Tawabini
B. S.
&
Laoui
T.
2017
A comparative study of raw and metal oxide impregnated carbon nanotubes for the adsorption of hexavalent chromium from aqueous solution
.
Bioinorganic Chemistry and Applications
2017
,
1624243
.
Rizzo
C.
,
Andrews
J. L.
,
Steed
J. W.
&
D'Anna
F.
2019
Carbohydrate-supramolecular gels: adsorbents for chromium (VI) removal from wastewater
.
Journal of Colloid and Interface Science
548
,
184
196
.
Rodiguez
M. H.
,
Yperman
J.
,
Carleer
R.
,
Maggen
J.
,
Dadi
D.
,
Gryglewicz
G.
,
Van der Bruggen
B.
,
Hernández
J. F.
&
Calvis
A. O.
2018
Adsorption of Ni (II) on spent coffee and coffee husk based activated carbon
.
Journal of Environmental Chemical Engineering
6
(
1
),
1161
1170
.
Sarici-Ozdemir
C.
2012
Adsorption and desorption kinetics behaviour of methylene blue onto activated carbon
.
Physicochemical Problems of Mineral Processing
48
(
2
),
441
454
.
Sethy
T. R.
&
Sahoo
P. K.
2019
Highly toxic Cr (VI) adsorption by (chitosan-g-PMMA)/silica bionanocomposite prepared via emulsifier-free emulsion polymerisation
.
International Journal of Biological Macromolecules
122
,
1184
1190
.
Shao
Z.
,
Huang
C.
,
Wu
Q.
,
Zhao
Y.
,
Xu
W.
,
Liu
Y.
,
Dang
J.
&
Hou
H.
2019
Ion exchange collaborating coordination substitution: More efficient Cr (VI) removal performance of a water-stable CuII-MOF material
.
Journal of Hazardous Materials
378
,
120719
.
Shi
Y.
,
Zhang
T.
,
Ren
H.
,
Kruse
A.
&
Cui
R.
2018
Polyethylene imine modified hydrochar adsorption for chromium (VI) and nickel (II) removal from aqueous solution
.
Bioresource Technology
247
,
370
379
.
Siddiqui
M. N.
,
Ali
I.
,
Asim
M.
&
Chanbasha
B.
2020
Quick removal of nickel metal ions in water using asphalt-based porous carbon
.
Journal of Molecular Liquids
308
,
113078
.
Siuki
A. K.
,
Shahidi
A.
,
Taherian
P.
&
Zeraatkar
Z.
2021
Comparing natural and mineral adsorbents in removing chromium from aquatic environment
.
Ain Shams Engineering Journal
12
(
3
),
2593
2601
.
Sivakumar
D.
,
Nouri
J.
,
Modhini
T.
&
Deepalakshmi
K.
2018
Nickel removal from electroplating industry wastewater: A bamboo activated carbon
.
Global Journal of Environmental Science and Management
4
(
3
),
325
338
.
Sohail
I.
,
Bhatti
I. A.
,
Ashar
A.
,
Sarim
F. M.
,
Mohsin
M.
,
Naveed
R.
,
Yasir
M.
,
Iqbal
M.
&
Nazir
A.
2020
Polyamidoamine (PAMAM) dendrimers synthesis, characterization and adsorptive removal of nickel ions from aqueous solution
.
Journal of Materials Research and Technology
9
(
1
),
498
506
.
Sun
L.
,
Wang
J.
,
Wu
J.
,
Wang
T.
,
Du
Y.
,
Li
Y.
&
Li
H.
2019
Constructing nanostructured silicates on diatomite for Pb (II) and Cd (II) removal
.
Journal of Materials Science
54
(
9
),
6882
6894
.
Veloso
C.
,
Filippov
L.
,
Filippova
I.
,
Ouvrard
S.
&
Araujo
A.
2020
Adsorption of polymers onto iron oxides: Equilibrium isotherms
.
Journal of Materials Research and Technology
9
(
1
),
779
788
.
Vilela
P. B.
,
Dalalibera
A.
,
Duminelli
E. C.
,
Becegato
V. A.
&
Paulino
A. T.
2019
Adsorption and removal of chromium (VI) contained in aqueous solutions using a chitosan-based hydrogel
.
Environmental Science and Pollution Research
26
(
28
),
28481
28489
.
Wu
Y.
,
Cha
L.
,
Fan
Y.
,
Fang
P.
,
Ming
Z.
&
Sha
H.
2017
Activated biochar prepared by pomelo peel using H 3 PO 4 for the adsorption of hexavalent chromium: performance and mechanism
.
Water, Air, & Soil Pollution
228
(
10
),
1
13
.
Xavier
A. L. P.
,
Adarme
O. F. H.
,
Furtado
L. M.
,
Ferreira
G. M. D.
,
da Silva
L. H. M.
,
Gil
L. F.
&
Gurgel
L. V. A.
2018
Modeling adsorption of copper (II), cobalt (II) and nickel (II) metal ions from aqueous solution onto a new carboxylated sugarcane bagasse. Part II: Optimization of monocomponent fixed-bed column adsorption
.
Journal of Colloid and Interface Science
516
,
431
445
.
Yu
J.
,
Zhang
J.
,
Song
S.
,
Liu
H.
,
Guo
Z.
&
Zhang
C.
2019
Removal of Ni (II) from aqueous solutions using activated carbon with manganese formate hydrate in-situ modification
.
Colloids and Surfaces A: Physicochemical and Engineering Aspects
560
,
84
91
.
Zhang
L.
,
Lu
H.
,
Yu
J.
,
Fan
Y.
,
Yang
Y.
,
Ma
J.
&
Wang
Z.
2018
Synthesis of lignocellulose-based composite hydrogel as a novel biosorbent for Cu2+ removal
.
Cellulose
25
(
12
),
7315
7328
.
Zhang
B.
,
Wu
Y.
&
Fan
Y.
2019
Synthesis of novel magnetic NiFe 2 O 4 nanocomposite grafted chitosan and the adsorption mechanism of Cr (VI)
.
Journal of Inorganic and Organometallic Polymers and Materials
29
(
1
),
290
301
.
Zhou
Z.
,
Kong
D.
,
Zhu
H.
,
Wang
N.
,
Wang
Z.
,
Wang
Q.
,
Liu
W.
,
Li
Q.
,
Zhang
W.
&
Ren
Z.
2018
Preparation and adsorption characteristics of an ion-imprinted polymer for fast removal of Ni (II) ions from aqueous solution
.
Journal of Hazardous Materials
341
,
355
364
.
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/).