Next Article in Journal
Co-Removal Effect and Mechanism of Cr(VI) and Cd(II) by Biochar-Supported Sulfide-Modified Nanoscale Zero-Valent Iron in a Binary System
Next Article in Special Issue
Sorption Preconcentration and Analytical Determination of Cu, Zr and Hf in Waste Samarium–Cobalt Magnet Samples
Previous Article in Journal
Annellated 1,3,4,2-Triazaphospholenes-Simple Modular Synthesis and a First Exploration of Ligand Properties
Previous Article in Special Issue
Microwave Irradiation and Glutamic Acid-Assisted Phytotreatment of Textile and Surgical Industrial Wastewater by Sorghum
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effective Removal of Methylene Blue from Simulated Wastewater Using ZnO-Chitosan Nanocomposites: Optimization, Kinetics, and Isotherm Studies

by
Zakariyya Uba Zango
1,*,
John Ojur Dennis
2,
A. I. Aljameel
3,*,
Fahad Usman
4,
Mohammed Khalil Mohammed Ali
3,
Bashir Abubakar Abdulkadir
2,
Saja Algessair
3,
Osamah A. Aldaghri
3 and
Khalid Hassan Ibnaouf
3
1
Department of Chemistry, Al-Qalam University Katsina, Katsina 2137, PMB, Nigeria
2
Department of Fundamental and Applied Science, Universiti Teknologi PETRONAS, Seri Iskandar 32610, Perak, Malaysia
3
Department of Physics, College of Science, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 13318, Saudi Arabia
4
Department of Physics, Al-Qalam University Katsina, Katsina 2137, PMB, Nigeria
*
Authors to whom correspondence should be addressed.
Submission received: 22 June 2022 / Revised: 12 July 2022 / Accepted: 22 July 2022 / Published: 25 July 2022

Abstract

:
Successful synthesis of ZnO-chitosan nanocomposites was conducted for the removal of methylene blue from an aqueous medium. Remarkable performance of the nanocomposites was demonstrated for the effective uptake of the dye, thereby achieving 83.77, 93.78 and 97.93 mg g−1 for the chitosan, 5 wt.% ZnO-Chitosan and 10 wt.% ZnO-Chitosan, respectively. The corresponding adsorption efficiency was 88.77, 93.78 and 97.95 for the chitosan, 5 wt.% ZnO-Chitosan and 10 wt.% ZnO-Chitosan, respectively. Upon regeneration, good reusability of the nanocomposites was manifested for the continuous removal of the dye up to six consecutive cycles. The adsorption process was kinetically described by a pseudo-first order model, while the isotherms were best fitted by the Langmuir model.

1. Introduction

Chitosan is a chitin derivative obtained by treating crustacean shells with sodium hydroxide through the acetylation process [1]. Having the presence of three types of functional groups—amino/acetamido group, primary and secondary hydroxyl groups on its backbone—chitosan is widely used as modifying agents in polymer composites [2,3]. It has shown great potential in industrial applications due to its non-toxic nature, biodegradability, and biocompatibility [4,5]. It has affinity for various organic and inorganic materials due to its abundant reaction sites. It can adsorb and chelate metal ions [6], and interact with bioactive molecules [7]. Because of its porosity, biocompatibility, biodegradability, non-toxicity, and robustness, it has versatile application in the preparations of novel chitosan-based materials for various applications such as pharmaceutical, nutraceutical [8], biomedical [9,10], fertilizer delivery [11], CO2 capture [5], catalysis [12] and wastewater remediations [13,14].
However, chitosan in its pure form is not as effective as its modified form. It is plagued by solubility deficiency, especially under acidic environments, which affects its mechanical properties for water related applications [15,16]. Thus, surface modification has been put forward to improve its structural properties. To increase the number of exposed active sites, various chemical or physical modifications have been adopted. Carbone nanotubes (CNTs) have been employed for the enhancement of the thermal stability of chitosan membranes and its reinforcement [17,18]. The use of the clay minerals such as smectite and sepiolite has been shown to improve its mechanical properties and resistance to aqueous media [19]. Similarly, the composites of chitosan with alginate have shown good chemical, thermal and moisture resistance [4]. Chitosan/halloysite composite membranes with excellent physico-chemical and thermal properties were also discovered [20]. Surfactant-modified chitosan (SMCS) beads have recently been reported for effective adsorption of heavy metals from the aqueous medium [21].
Recently, surface coating of the chitosan with metal oxide nanoparticles has gained the recognition of researchers for adsorption, catalysis, degradation, and biomedical applications [22,23]. A facile and greener synthesis of chitosan-FeO nanocomposite was reported by Bharathi et al. (2019). The good structural and absorption properties of the biopolymer material were ascertained, and its potential antibacterial activity was evaluated [9]. Neeraj et al. (2016) investigated the elimination of arsenic from the aqueous solution using a nanocomposite of chitosan coated with iron-oxide. The porosity of the material was confirmed with a cumulative pore volume of 0.0362 m3/g and pore diameter of 32.46 nm. Thus, the maximum monolayer adsorption capacity of the material was 267.2 mg/g [24]. Synthesis and characterizations of nano hydrogel beads of chitosan/agar/SiO2 composites were also reported for effective pharmaceutical adsorption from environmental waters, achieving over 99% adsorption of naproxen within 15 min of the batch adsorption process [25].
Rapid detection of dyes in various water bodies has been an environmental phenomenon over the years [26,27]. In this research, we aimed at synthesizing water resistant ZnO-chitosan nanocomposites for the efficient adsorption of dyes from the aqueous medium. Methylene blue was chosen as model dye due to its large consumption by textile, leather, tannery, paper, and pulp processing industries [28]. Thus, the adsorption parameters, kinetics, isotherms, regeneration, and reusability of the adsorbent materials were evaluated for optimum uptake of the dye from the aqueous medium.

2. Results

2.1. Materials

Chitosan (85% deacetylated, Molecular weight of 400,000 Da) is purchased from Golden-Shell Biochemical (Yuhuan, Zhejiang, China). The crystal violet (Molecular Formula C25N3H30Cl, MW 407.979 g/mol and Density: 1.19 g/cm3) and the zinc oxide nano powder of particle size 50–100 nm and 98% purity were purchased from Sigma-Aldrich (St. Louis, MO, USA). Acetic acid and ethanol were purchased from Avantis Laboratory, Perak, Malaysia. All other chemicals not mentioned here were of high purity and of analytical reagent (AR) grade and were used as received. Double distilled water was used throughout the study.

2.2. Synthesis of ZnO-Chitosan Nanocomposite

The nanocomposites were synthesized by dissolving 10 mg of ZnO Nano powder in 90 mL ethanol with constant stirring for 15 min. A 50 mg of Chitosan was dispersed in 20 mL of 0.1 M acetic acid and stirred for 15 min. The 10 mL ZnO solution was gradually added to the dispersed chitosan. The mixture was continuously stirred for 2 h, and the resulting mixture was centrifuged at 4000 rpm, filtered and washed thoroughly with ethanol and double distilled water. The obtained residue was dried overnight in an oven at 100 °C and stored in vacuum. It was labeled as 10 wt.% ZnO-chitosan. For the preparation of 5 wt.% ZnO-chitosan, the same method was employed; only 5 mL of the ZnO solution was used.

2.3. Characterizations

The materials were examined by Scanning Electron Microscopy (SEM) (SU8020, Hitachi, Tokyo, Japan) for the morphology analysis. The samples were coated with gold by a Polaran (SC 515) sputter coater to make it electrically conductive. N2 adsorption–desorption of the samples were determined by a TriStar II (3020) Micromeritics porosity analyzer, Norcross, GA, USA. The sample was heated at 60 °C, and the analysis was conducted under a nitrogen stream with adsorption–desorption isotherms at 77 K.

2.4. Preparation of Methylene Blue Solution

The dye stock solution was prepared by dissolving 1000 mg of the methylene blue in 100 mL distilled water. The solution was transferred to a 1000 mL volumetric flask, and distilled water was added to the mark, making a concentration of 1000 mg L−1, which was kept in a refrigerator prior to the adsorption experiments. The solutions of different concentrations used in various experiments were obtained by a dilution of the stock solutions.

2.5. Batch Adsorption Experiment

The adsorption behavior of chitosan and the nanocomposites were evaluated for the methylene blue adsorption. Batch experiments were conducted using 50 mL of 100 mg L−1 of the dye solutions in 100 mL conical flasks with the adsorbent of 6 g/L. The conical flasks were placed in a thermostatic shaker at 250 rpm and 30 °C. The absorbance of the solutions was analyzed by using a UV–vis spectrophotometer (Varian CARY 50 probe) at 665 nm.
The optimum time can be obtained from the plot of adsorption capacity versus time using the formula:
q t = ( C o C t ) V w
In addition, the equilibrium time for the adsorption of the dye is obtained using the formula:
q e = ( C o C e ) V w
and the removal efficiency (%R) is calculated from the formula:
( % R ) = C o C e C o × 100
where C o and C e are initial and equilibrium concentrations of the dye (mg L−1), respectively.

2.6. Adsorption Kinetics

The data for the batch adsorption studied were evaluated by pseudo-first order, pseudo-second order, and intra-particle diffusion described by the Equations (5)–(7), respectively, for the determination of the best kinetic model:
ln ( q e q t ) = ln q e k 1 t
t / q t = 1 k 2 q e 2 + t q e
q t = k p t 1 / 2 + C
where q e and q t represent the amounts of dye adsorbed (mg/g) at equilibrium and time t, respectively. k 1 (1/min) and k 2 (g mg−1 min−1) are the rate constants of the pseudo first-order and pseudo second-order adsorption kinetic model, respectively. k p is the intraparticle diffusion rate constant and C being constant.

2.7. Adsorption Isotherms

To evaluate the isotherms fitting on the adsorption data, the models of Langmuir, Freundlich, and Temkin were employed as represented by Equations (7)–(9), respectively:
C e q e = 1 K L q m + 1 q m C e
log q e = log K F + 1 n log C e
q e = R T b T ln A T + ( R T b T ) ln C e
where q m (mg g−1) is the Langmuir adsorption capacity, K L is the Langmuir constant (L mg−1), K F is the Freundlich constant (L mg−1), 1/n represents the adsorption intensity, b T (kJ mol−1) represents the Temkin constant which relates to the heat of adsorption, and A T is the equilibrium binding constant corresponding to the maximum binding energy (L g−1). T is the absolute temperature (K), and R (8.314 J mol−1 K−1) is the Universal gas constant.
The favorability of the adsorption process was determined from the Langmuir constant given by Equation (10):
R L = 1 1 + C 0 K L
When the value of R L is less than unity, the adsorption is favorable and when greater than unity is considered as unfavorable. In addition, when R L is 0, the adsorption is irreversible and linear when it is unity.
Moreover, these models were statistically analyzed using regression to determine the coefficient of determination (R2), root mean square error (RMSE), and Akaike information criterion (AIC) to assess the model performance, using Equations (11)–(13):
R 2 = 1 ( q e   e x p q e   c a l ) 2 ( q e   e x p ) 2
R M S E = n = 1 i ( q e   e x p q e   m o d e l ) 2
A I C = n ln ( S S E n ) + 2 n p + 2 n p ( n p + 1 ) n ( n p + 1 )
where q e   e x p and q e   m o d e l represent experimental and model adsorption capacity, n is the number of observations, and p denotes the number of parameters. SSE is the sum of the square errors obtained. Higher R2 value indicates better linearity of the models while smaller RMSE and AIC indicate better fitting of the model.

2.8. Effect of Adsorbent Dosage

The adsorbents dosage was varied from 1–6 g L−1, and, using the initial dye concentration of 100 mg L−1 and pH at 4, the adsorption experiment was conducted. The removal efficiency (%R) was plotted against adsorbent dosage

2.9. Effect of pH

Solution pH affects the adsorption process by affecting both aqueous chemistry and surface binding sites of the adsorbent. In this work, the pH range was studied from 2–12 using an initial dye concentration of 100 mg L−1. The pH was adjusted with 0.1 M HCl or NaOH and measured with pH-meter model HI 8014, Hanna Instruments (Padua, Italy).

2.10. Adsorbent Regeneration and Reusability Test

For the adsorbent regeneration and reusability, an adsorption experiment was conducted with an initial dye concentration of 100 mg L−1 and adsorbent dosage of 6 g L−1 at room temperature and stirring rate of 250 rpm. After the adsorption, the adsorbent was centrifuged, filtered, and washed with 0.1 M NaOH solution and then several times with distilled water to remove all the traces of the dye [29,30]. The experiments were repeated several times using the same procedure.

3. Discussion

3.1. Characterizations

The characteristic absorption band of the pristine chitosan and nanocomposites was studied using UV-Visible absorption spectroscopy. As shown in Figure 1, the composites have exhibited strong absorption band at around 350–380 nm, compared to the pristine chitosan with the corresponding calculated band gap of 3.62 eV. This corresponds to the band gap of the pristine ZnO nanoparticles having a value of 3.37 eV with an absorption maximum of 368 nm as previously reported by Rao et al. [31]. Thus, the spectra confirmed the dispersion of the ZnO nanoparticles on the surface of the chitosan and the good adsorption properties of the ZnO-chitosan nanocomposites.
The morphology of the prepared composites is shown the SEM images as depicted in Figure 2. It indicated the spherical or elliptical shape of the ZnO nanoparticles on the surface of the chitosan. Previous studies have indicated that the presence of metal nanoparticles enhanced the adsorption performance of the pristine chitosan, such as the work of Rahmi et al. (2019) for the adsorption of Hg (II) and Cd (II) onto highly crystalline Fe2O3@chitosan nanocomposite [16]. Most recent is the finding of Moradi et al. for the naproxen adsorption onto Chitosan/agar/SiO2 nano hydrogels [25]. The average particle size was 3.46 and 3.51 nm for the 5 wt.% and 10 wt.% ZnO-chitosan, respectively, as depicted in Figure 3.
The porosity of the adsorbent materials is vital for adsorption studies as the pores in the adsorbents offer more adsorption sites for the guest molecules. Adsorbents having higher N2 adsorption–desorption isotherms often presented higher adsorption sites for the host molecules. The specific Brunauer–Emmette–Teller (BET) surface area of the pristine chitosan was 4.22 m2 g. However, for the nano composites, BET surface area of 45.70 and 49.21 m2 g−1 were recorded for 5 wt.% ZnO-chitosan and 10 wt.% ZnO-chitosan samples, respectively, as shown in Table 1. The corresponding values for pore volumes and the average pore diameter according to the Barrett–Joyner–Halenda (BJH) were highlighted in Table 1. The higher surface area of the nanocomposites revealed their good efficiency for the uptake of the dye molecules. Previous reports have also indicated that the synergetic effect of metal-oxide nanoparticles affects the overall surface area of pristine chitosan [23].

3.2. Effect of Contact Time

Contact time is paramount variable in adsorption processes. Thus, the effect of contact time on the dye adsorption onto the pristine chitosan and the nanocomposites were investigated. Figure 4 has shown the increase in adsorption capacity ( q t mg g−1) with the contact time for both chitosan and nanocomposites. This might be attributed to the diffusion of dye molecules from the surface of the solution onto the surface of the adsorbents [32]. However, with time, the adsorption capacity becomes moderate, probably due to the migration of dye molecules to inner pores of the adsorbents. The equilibrium was achieved at 160 min with equilibrium adsorption capacity of 93.78 mg g−1 and 97.93 mg g−1 for 5 wt.% ZnO-Chitosan and 10 wt.% ZnO-Chitosan, respectively. In comparison, the pristine chitosan has a q e value of 83.77 mg g−1. These results indicated an improvement in the adsorption capacity of the nanocomposites when compared with the chitosan. A similar observation was reported when the chitosan was modified with PVA/TiO2 [33].

3.3. Adsorbent Dosage

A promising adsorbent material must be able to remove considerable amounts of adsorbate at low doses. This feature is paramount to reduce operational costs and minimize the risks of secondary pollution [34,35]. From the plot of Figure 5, the removal efficiency (%R) against adsorbent dosage (g L−1), it was shown that adsorption capacity increased with the amount of the adsorbent. This was due to the increase in number of active sites. Thus, adsorbent dosage of 6 g L−1 is required to remove the dye concentration of 100 mg L−1, achieving the higher adsorption efficiency of 87.54, 96.39, and 99.95% for the chitosan, 5 wt.% ZnO-chitosan, and 10 wt.% ZnO-chitosan composites, respectively [36,37].

3.4. Effect of Dye Concentration

The adsorption of the dye onto adsorbents was studied at different initial concentrations. From Figure 6, the adsorption efficiency decreased for all the adsorbents when the initial concentration was changed from 50 to 250 mg L−1. For the pristine chitosan, the adsorption efficiency drastically decreased from 96.17 to 42.21%, while, for the nanocomposites, the adsorption efficiency decreased from 99.53 to 58. 83% and from 99.95 to 61.51% for the 5 wt.% ZnO-chitosan and 10 wt.% ZnO-chitosan, respectively. The decrease in the adsorption efficiency at higher concentration was due to the limited number of the adsorption sites available for the uptake of the dye. As the adsorption sites became saturated, no more adsorption occurred. However, the adsorption capacity ( q e mg g−1) increased as the initial concentration of the dye was increased due to the mass driving force that enables the transfer of the dye molecules to the active sites of the adsorbents.

3.5. Influence of pH on Adsorption

The adsorption capacity gradually increased at the acidic pH until it reached the peak at the pH of 8. It then started to decline at the alkaline pH (Figure 7). The increase in the adsorption capacity observed at the lower pH was resulted from the attraction of the amino group on the surface of the chitosan for the hydroxonium ion on the surface of the solution [38]. Also at the basic pH above 7, deprotonation of the active sites of the surface of the adsorbent occurred, thus repulsed with the OH on the surface of the [39,40]. The presence of excess H+ and OH in the acidic or basic solution competes with the dye for the adsorption sites, which resulted in lower adsorption capacity.

3.6. Kinetics of Adsorption

The mechanism and the rate controlling step of the adsorption process was evaluated by the kinetics models of pseudo-first order and pseudo-second order intra-particle diffusion. Among models, the pseudo-first order has shown the best calculated adsorption capacities of 85.308, 98.397 and 103.048 mg g−1 for the chitosan, 5 wt.% ZnO-chitosan and 10 wt.% ZnO-chitosan, respectively. This has been in good agreement with the experimental results. Additionally, the fitting data of the model have shown best R2 values 0.998, 0.995 and 0.959 for the chitosan, 5 wt.% ZnO-chitosan and 10 wt.% ZnO-chitosan, respectively. Similarly, the statistical values for the linear regression analysis of MSE, RMSE and AIC of the model were also in good agreement with the finding as highlighted in Table 2. Thus, the adsorption of the dye onto the pristine chitosan and the nanocomposites best described the abundant adsorption sites as depicted by the improvement in the BET surface area upon the dispersion of the ZnO nanoparticle on the surface of the chitosan.

3.7. Isotherms of Adsorption

The isotherms studies were used to describe the interaction between the dye and the adsorbents when the equilibrium is attained. Of the models studied (Table 3), Langmuir fitting was the most consistent for the adsorption data according to the obtained R2 values and the statistical regression analysis as highlighted in Table 4. The Langmuir adsorption capacity ( q m ) was 68.077, 87.471 and 90.976 mg/g for the chitosan, 5 wt.% ZnO-Chitosan and 10 wt.% ZnO-chitosan, respectively, indicating the monolayer formation and the good adsorption capacity of the adsorbents [27]. Similarly, the corresponding RL values were 0.043, 0.224 and 0.019 for the chitosan, 5 wt.% ZnO-Chitosan and 10 wt.% ZnO-chitosan, respectively, signifying the favorability of the adsorption. Thus, the overall adsorption process is said to occur via monolayer formation. Previously, Zhang et al. have reported similar observations for the adsorption of cesium chitosan-vermiculite composite [41].

3.8. Adsorbent Regeneration and Reusability

Results from regeneration and reusability studies indicated that both chitosan and the nanocomposites can withstand the dye adsorption for number of repeated usages with good efficiency. From Figure 8, the adsorption efficiency of the chitosan dropped from 84.54% to 50.93%, whereas it dropped from 93.39% to 54.04% and 99.95% to 59.05% for the 5 wt.% ZnO-chitosan and 10 wt.% ZnO-chitosan, for the 1st and 6th cycles, respectively. This reaffirmed that both chitosan and composites are good adsorbents for the dye removal over a repeated number of adsorption cycles.
Literature studies have also restated the good performance of the nanocomposites in comparison to other chitosan and composites reported. The efficiency of the ZnO-chitosan nanocomposites could be deduced from the higher adsorption capacities of the materials and the shorter equilibration time than most of the adsorbents previously employed. Thus, the relevance of this work in the field of pollutants’ remediation from environmental waters.

4. Conclusions

ZnO nanoparticles were successfully dispersed on chitosan, forming 5 wt.% ZnO-Chitosan and 10 wt./% ZnO-Chitosan nanocomposites. The UV-visible analysis has shown the good adsorption of the composites, while the SEM analysis described the surface morphology of the nanocomposites with the spherical or elliptical shape of the ZnO nanoparticles, indicating the formation of the ZnO-chitosan nanocomposites. The N2 adsorption–desorption analysis revealed the good porosities of the composites for the uptake of the guest molecules. The adsorption studies for the removal of methylene blue from the aqueous medium demonstrated efficiency of the nanocomposites and its higher adsorption capacity compared to the pristine chitosan with the equilibrium attained within 160 min. The adsorption capacities were 83.77, 93.78 and 97.93 mg g−1 for the chitosan, 5 wt.% ZnO-Chitosan and 10 wt.% ZnO-Chitosan, respectively. The good adsorbent properties of the materials were demonstrated for the efficient adsorption of the methylene blue up to six cycles, achieving remarkable adsorption efficiency. The kinetics and isotherms were governed by pseudo-first order and Langmuir model, respectively. Thus, the nanocomposites can be employed as good adsorbents for pollutants removal from the environmental waters. However, for real sample application, the multi-component adsorption system should be employed using column technology and a suitable experimental design model.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/molecules27154746/s1, Figure S1: Pseudo-first order model plot for methylene blue adsorption onto chitosan and the nanocomposites; Figure S2: Pseudo-second order model plot for methylene blue adsorption onto chitosan and the nanocomposites; Figure S3: Intraparticle adsorption model plot for methylene blue adsorption onto chitosan and the nanocomposites; Figure S4: Langmuir isotherm model plot for methylene blue adsorption onto chitosan and the nanocomposites; Figure S5: Freundlich isotherm model plot for methylene blue adsorption onto chitosan and the nanocomposites; Figure S6: Temkin isotherm model plot for methylene blue adsorption onto chitosan and the nanocomposites.

Author Contributions

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

Funding

This work was funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud University RG-21-09-50.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in Supplementary Material.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University for funding this work through Research Group no. RG-21-09-50.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Chang, M.Y. Adsorption of tannic acid, humic acid, and dyes fromwater using the composite of chitosan and activated clay. J. Colloid Interface Sci. 2004, 278, 18–25. [Google Scholar] [CrossRef]
  2. Don, T.M.; King, C.F.; Chiu, W.Y. Synthesis and properties of chitosanmodified poly (vinyl acetate). J. Appl. Polym. Sci. 2002, 86, 3057–3063. [Google Scholar] [CrossRef]
  3. Babaee, M.; Garavand, F.; Rehman, A.; Jafarazadeh, S.; Amini, E.; Cacciotti, I. Biodegradability, physical, mechanical and antimicrobial attributes of starch nanocomposites containing chitosan nanoparticles. Int. J. Biol. Macromol. 2022, 195, 49–58. [Google Scholar] [CrossRef]
  4. Li, Y.; Sun, S.; Gao, P.; Zhang, M.; Fan, C.; Lu, Q.; Li, C.; Chen, C.; Lin, B.; Jiang, Y. A tough chitosan-alginate porous hydrogel prepared by simple foaming method. J. Solid State Chem. 2021, 294, 121797. [Google Scholar] [CrossRef]
  5. Wang, K.; Wang, H.; Pan, S.; Fu, C.; Chang, Y.; Li, H.; Yang, X.; Qi, Z. Evaluation of New Film Based on Chitosan/Gold Nanocomposites on Antibacterial Property and Wound-Healing Efficacy. Adv. Mater. Sci. Eng. 2020, 2020, 6212540. [Google Scholar] [CrossRef]
  6. Rosli, N.; Yahya, W.Z.N.; Wirzal, M.D.H. Crosslinked chitosan/poly(vinyl alcohol) nanofibers functionalized by ionic liquid for heavy metal ions removal. Int. J. Biol. Macromol. 2022, 195, 132–141. [Google Scholar] [CrossRef]
  7. Namasivayam, S.K.R.; Samrat, K.; Debnath, S.; Jayaprakash, C. Biocompatible chitosan nanoparticles incorporated bacteriocin (CSNps-B) preparation for the controlled release and improved anti-bacterial activity against food borne pathogenic bacteria Listeria monocytogenes. Res. J. Pharm. Biol. Chem. Sci. 2015, 6, 625–631. [Google Scholar]
  8. Li, B.; Elango, J. Recent Advancement of Molecular Structure and Biomaterial Function of Chitosan from Marine Organisms for Pharmaceutical and Nutraceutical Application. Appl. Sci. 2020, 10, 4719. [Google Scholar] [CrossRef]
  9. Bharathi, D.; Ranjithkumar, R.; Vasantharaj, S.; Chandarshekar, B.; Bhuvaneshwari, V. Synthesis and characterization of chitosan/iron oxide nanocomposite for biomedical applications. Int. J. Biol. Macromol. 2019, 132, 880–887. [Google Scholar] [CrossRef]
  10. Wang, K.; Pan, S.; Qi, Z.; Xia, P.; Xu, H.; Kong, W.; Li, H.; Xue, P.; Yang, X.; Fu, C. Recent Advances in Chitosan-Based Metal Nanocomposites for Wound Healing Applications. Adv. Mater. Sci. Eng. 2020, 2020, 3827912. [Google Scholar] [CrossRef]
  11. Rodrigues, B.; Britti, F.; Pereira, S. Chitosan-Montmorillonite microspheres: A sustainable fertilizer delivery system. Carbohydr. Polym. 2015, 127, 340–346. [Google Scholar] [CrossRef]
  12. Dhakshinamoorthy, A.; Jacob, M.; Sakthi, N.; Varalakshmi, P. Pristine and modified chitosan as solid catalysts for catalysis and biodiesel production: A minireview. Int. J. Biol. Macromol. 2021, 167, 807–833. [Google Scholar] [CrossRef] [PubMed]
  13. Popuri, S.R.; Vijaya, Y.; Boddu, V.M.; Abburi, K. Adsorptive removal of copper and nickel ions from water using chitosan coated PVC beads. Bioresour. Technol. 2009, 100, 194–199. [Google Scholar] [CrossRef]
  14. Vakili, M.; Rafatullah, M.; Salamatinia, B.; Abdullah, A.Z.; Ibrahim, M.H.; Tan, K. Application of chitosan and its derivatives as adsorbents for dye removal from water and wastewater: A review. Carbohydr. Polym. 2014, 113, 115–130. [Google Scholar] [CrossRef]
  15. Kyzas, G.Z.; Bikiaris, D.N. Recent modifications of chitosan for adsorption applications: A critical and systematic review. Mar. Drugs 2015, 13, 312–337. [Google Scholar] [CrossRef] [PubMed]
  16. Rahmi; Fathurrahmi; Lelifajri; Purnamawati, F.; Sembiring, R. Preparation of Magnetic Chitosan Beads for Heavy Metal Ions Removal from Water. IOP Conf. Ser. Earth Environ. Sci. 2019, 276, 012004. [Google Scholar] [CrossRef] [Green Version]
  17. Mousavi, S.R.; Asghari, M.; Mahmoodi, N.M. Chitosan-wrapped multiwalled carbon nanotube as filler within PEBA thin film nanocomposite (TFN) membrane to improve dye removal. Carbohydr. Polym. 2020, 237, 116128. [Google Scholar] [CrossRef] [PubMed]
  18. Borgohain, R.; Jain, N.; Prasad, B.; Mandal, B.; Su, B. Carboxymethyl chitosan/carbon nanotubes mixed matrix membranes for CO2 separation. React. Funct. Polym. 2019, 143, 104331. [Google Scholar] [CrossRef]
  19. Padilla-Ortega, E.; Darder, M.; Aranda, P.; Figueredo Gouveia, R.; Leyva-Ramos, R.; Ruiz-Hitzky, E. Ultrasound assisted preparation of chitosan–vermiculite bionanocomposite foams for cadmium uptake. Appl. Clay Sci. 2016, 130, 40–49. [Google Scholar] [CrossRef]
  20. De Silva, R.T.; Pasbakhsh, P.; Goh, K.L.; Chai, S.-P.; Ismaild, H. Physico-chemical characterisation of chitosan/halloysite composite membranes. Polym. Test. 2013, 32, 265–271. [Google Scholar] [CrossRef]
  21. Pal, P.; Pal, A. Surfactant-modified chitosan beads for cadmium ion adsorption. Int. J. Biol. Macromol. 2016, 104, 1548–1555. [Google Scholar] [CrossRef]
  22. Senthilkumar, P.; Yaswant, G.; Kavitha, S.; Chandramohan, E.; Kowsalya, G.; Vijay, R.; Sudhagar, B.; Kumar, D.S.R.S. Preparation and characterization of hybrid chitosan-silver nanoparticles (Chi-Ag NPs); A potential antibacterial agent. Int. J. Biol. Macromol. 2019, 141, 290–297. [Google Scholar] [CrossRef]
  23. Rani, M.; Rachna; Shanker, U. Metal oxide-chitosan based nanocomposites for efficient degradation of carcinogenic PAHs. J. Environ. Chem. Eng. 2020, 8, 103810. [Google Scholar] [CrossRef]
  24. Gerard, N.; Raghunandan, S.K.; Kumar, P.S.; Cabana, H.; Vaidyanathan, V.K. Adsorptive potential of dispersible chitosan coated iron-oxide nanocomposites toward the elimination of arsenic from aqueous solution. Process Saf. Environ. Prot. 2016, 104, 185–195. [Google Scholar] [CrossRef]
  25. Moradi, O.; Mhdavi, S.; Sedaghat, S. Synthesis and Characterization of Chitosan/Agar/SiO2 Nano Hydrogels for Removal of Amoxicillin and of Naproxen from Pharmaceutical Contaminants. Res. Sq. 2021, 1–23. [Google Scholar] [CrossRef]
  26. Imam, S.S.; Muhammad, A.I.; Babamale, H.F.; Zango, Z.U. Removal of Orange G Dye from Aqueous Solution by Adsorption: A Short Review. J. Environ. Treat. Tech. 2021, 9, 318–327. [Google Scholar] [CrossRef]
  27. Zango, Z.U.; Dahiru, M.; Haruna, M.A. Adsorption kinetics, isotherms and thermodynamics for malachite green and methylene blue removal in water using low-cost banana peel biosorbent. Res. J. Chem. Environ. 2020, 24, 287–298. [Google Scholar]
  28. Zango, Z.U.; Imam, S.S. Microcrystalline Cellulose from Groundnut Shell as Potential Adsorbent of Crystal Violet and Methylene Blue. Kinetics, Isotherms and Thermodynamic Studies. Chem. Chem. Technol. 2020, 14, 563–571. [Google Scholar] [CrossRef]
  29. Isiyaka, H.A.; Jumbri, K.; Sambudi, N.S.; Lim, J.W.; Saad, B.; Ramli, A.; Zango, Z.U. Experimental and Modeling of Dicamba Adsorption in Aqueous Medium Using MIL-101(Cr) Metal-Organic Framework. Processes 2021, 9, 419. [Google Scholar] [CrossRef]
  30. Zhai, L.; Bai, Z.; Zhu, Y.; Wang, B.; Luo, W. Fabrication of chitosan microspheres for efficient adsorption of methyl orange. Chin. J. Chem. Eng. 2018, 26, 657–666. [Google Scholar] [CrossRef]
  31. Rao, U.S.; Srinivas, G.; Rao, T.P. Influence of precursors on morphology and spectroscopic properties of ZnO Nanoparticles. Procedia Mater. Sci. 2015, 10, 90–96. [Google Scholar] [CrossRef] [Green Version]
  32. Garba, Z.N.; Tanimu, A.; Zango, Z.U. Borassus aethiopum shell-based activated carbon as efficient adsorbent for carbofuran. Bull. Chem. Soc. Ethiop. 2019, 33, 425–436. [Google Scholar] [CrossRef]
  33. Habiba, U.; Joo, T.C.; Siddique, T.A.; Salleh, A.; Ang, B.C.; Afifi, A.M. Effect of degree of deacetylation of chitosan on adsorption capacity and reusability of chitosan/polyvinyl alcohol/TiO2 nano composite. Int. J. Biol. Macromol. 2017, 104, 1133–1142. [Google Scholar] [CrossRef]
  34. Fosso-Kankeu, E.; Mittal, H.; Mishra, S.B.; Mishra, A.K. Gum ghatti and acrylic acid based biodegradable hydrogels for the effective adsorption of cationic dyes. J. Ind. Eng. Chem. 2015, 22, 171–178. [Google Scholar] [CrossRef]
  35. Dahiru, M.; Zango, Z.U.; Haruna, M.A. Cationic Dyes Removal Using Low-Cost Banana Peel Biosorbent. Am. J. Mater. Sci. 2018, 8, 32–38. [Google Scholar] [CrossRef]
  36. Swayampakula, K.; Boddu, B.M.; Nadavala, K.A. Competitiveadsorption of Cu (II): Co (II) and Ni (II) from their binary and tertiary aqueoussolutions using chitosan-coated perlite beads as biosorbent. J. Hazard. Mater 2009, 170, 680–689. [Google Scholar] [CrossRef]
  37. Naseeruteen, F.; Hamid, N.S.A.; Suah, F.B.M.; Ngah, W.S.W.; Mehamod, F.S. Adsorption of malachite green from aqueous solution by using novel chitosan ionic liquid beads. Int. J. Biol. Macromol. 2018, 107, 1270–1277. [Google Scholar] [CrossRef]
  38. Fan, Z.; Xin, C.; Fenghuang, Y.J. High adsorption capability and selectivity of ZnO nanoparticles for dye removal. Colloids Surf. A Physicochem. Eng. Asp. 2016, 509, 474–483. [Google Scholar]
  39. Chen, L.; Wu, P.; Chen, M.; Lai, X.; Ahmed, Z.; Zhu, N.; Dang, Z.; Bi, Y.; Liu, T. Preparation and characterization of the eco-friendly chitosan/vermiculite biocomposite with excellent removal capacity for cadmium and lead. Appl. Clay Sci. 2018, 159, 74–82. [Google Scholar] [CrossRef]
  40. Zhang, F.; Song, W.J.; Lan, J. Effective removal of methyl blue by fine-structured strontium and barium phosphate nanorods. Appl. Surf. Sci. 2015, 326, 195–203. [Google Scholar] [CrossRef]
  41. Zhang, B.; Zhang, B.; Liu, X. Chitosan coated-porous low expansion vermiculite for efficient removal of cesium from radioactive wastewater. Environ. Chem. Ecotoxicol. 2021, 3, 182–196. [Google Scholar] [CrossRef]
  42. Uzun, I. Kinetics of the adsorption of reactive dyes by chitosan. Dyes Pigments 2006, 70, 76–83. [Google Scholar] [CrossRef]
  43. Arunachalam, K.D. Bio-adsorption of methylene blue dye using chitosan-extracted from Fenneropenaeus indicus shrimp shell waste. J. Aquac. Mar. Biol. 2021, 10, 146–150. [Google Scholar] [CrossRef]
  44. Shashikala, M.; Nagapadma, M.; Pinto, L.; Nambiar, S.N. Studies on The Removal of Methylene Blue Dye From Water Using Chitosan. Int. J. Dev. Res. 2013, 3, 40–44. [Google Scholar]
  45. Pietrelli, L.; Francolini, I.; Piozzi, A. Dyes Adsorption from Aqueous Solutions by Chitosan. Sep. Sci. Technol. 2015, 50, 1101–1107. [Google Scholar] [CrossRef] [Green Version]
  46. Murcia-Salvador, A.; Pellicer, J.A.; Fortea, M.I.; Gómez-López, V.M.; Rodríguez-López, M.I.; Núñez-Delicado, E.; Gabaldón, J.A. Adsorption of Direct Blue 78 using chitosan and cyclodextrins as adsorbents. Polymers 2019, 11, 1003. [Google Scholar] [CrossRef] [Green Version]
  47. Mohamed, H.S.; Soliman, N.K.; Moustafa, A.F.; Abdel-Gawad, O.F.; Taha, R.R.; Ahmed, S.A. Nano metal oxide impregnated Chitosan-4-nitroacetophenone for industrial dye removal. Int. J. Environ. Anal. Chem. 2021, 101, 1850–1877. [Google Scholar] [CrossRef]
  48. Zhu, H.; Zhang, M.; Liu, Y.; Zhang, L.; Han, R. Study of congo red adsorption onto chitosan coated magnetic iron oxide in batch mode. Desalin. Water Treat. 2012, 37, 46–54. [Google Scholar] [CrossRef]
  49. Tanhaei, B.; Ayati, A.; Lahtinen, M.; Sillanpää, M. Preparation and characterization of a novel chitosan/Al2O3/magnetite nanoparticles composite adsorbent for kinetic, thermodynamic and isotherm studies of Methyl Orange adsorption. Chem. Eng. J. 2015, 259, 1–10. [Google Scholar] [CrossRef]
  50. Ge, Y.; Zhao, X.; Xu, J.; Liu, J.; Yang, J.; Li, S. Recyclable magnetic chitosan microspheres with good ability of removing cationic dyes from aqueous solutions. Int. J. Biol. Macromol. 2021, 167, 1020–1029. [Google Scholar] [CrossRef]
Figure 1. UV-visible absorbance spectra of 5 wt.% ZnO-chitosan and 10 wt.% ZnO-chitosan nanocomposites.
Figure 1. UV-visible absorbance spectra of 5 wt.% ZnO-chitosan and 10 wt.% ZnO-chitosan nanocomposites.
Molecules 27 04746 g001
Figure 2. SEM of Chitosan (a) 5 wt.% ZnO-chitosan (b) 10 wt.% ZnO-chitosan (c) at 1 K and 5 K magnifications, respectively.
Figure 2. SEM of Chitosan (a) 5 wt.% ZnO-chitosan (b) 10 wt.% ZnO-chitosan (c) at 1 K and 5 K magnifications, respectively.
Molecules 27 04746 g002
Figure 3. Particle size distribution of ZnO in 5 wt.% ZnO-chitosan (a) and 10 wt.% ZnO-chitosan (b) nanocomposites.
Figure 3. Particle size distribution of ZnO in 5 wt.% ZnO-chitosan (a) and 10 wt.% ZnO-chitosan (b) nanocomposites.
Molecules 27 04746 g003
Figure 4. Effect of contact time for the dye adsorption onto chitosan and the nanocomposites.
Figure 4. Effect of contact time for the dye adsorption onto chitosan and the nanocomposites.
Molecules 27 04746 g004
Figure 5. Effects of adsorbent dosage for the dye adsorption onto the chitosan and the nanocomposite.
Figure 5. Effects of adsorbent dosage for the dye adsorption onto the chitosan and the nanocomposite.
Molecules 27 04746 g005
Figure 6. Effect of initial dye concentrations for the dye adsorption onto the chitosan and nanocomposites.
Figure 6. Effect of initial dye concentrations for the dye adsorption onto the chitosan and nanocomposites.
Molecules 27 04746 g006
Figure 7. Effects of pH for the dye adsorption onto the chitosan and the nanocomposites.
Figure 7. Effects of pH for the dye adsorption onto the chitosan and the nanocomposites.
Molecules 27 04746 g007
Figure 8. Reusability of chitosan and composites for the dye adsorption.
Figure 8. Reusability of chitosan and composites for the dye adsorption.
Molecules 27 04746 g008
Table 1. N2 adsorption–desorption properties of chitosan and the composites.
Table 1. N2 adsorption–desorption properties of chitosan and the composites.
PropertiesChitosan5 wt.% ZnO-Chitosan10 wt.% ZnO-Chitosan
BET surface area (m2 g−1)4.2245.7049.21
Pore volume (m3 g−1)0.0130.0250.025
Pore sizes (nm)4.154.624.62
Table 2. Kinetics models for the dye adsorption onto the chitosan and the nanocomposites.
Table 2. Kinetics models for the dye adsorption onto the chitosan and the nanocomposites.
ModelChitosan5 wt.% ZnO-Chitosan10 wt.% ZnO-Chitosan
qe (Experimental mg g−1)88.77193.78097.951
Pseudo-first order
qe (Calculated mg g−1)85.30898.397103.048
K 1 (min−1)0.0100.0120.013
R20.9980.9950.959
R2 adj0.9780.9830.943
MSE0.0200.0270.026
RMSE0.1410.1630.160
AIC−33.141−30.904−31.197
Pseudo-second order
qe (Calculated mg g−1)16.25027.65130.150
K 2 (g mg−1 min−1)0.0000.0000.000
R20.7070.7720.909
R2 adj0.6650.7390.896
MSE0.1720.0630.024
RMSE0.4140.2510.156
AIC−14.119−23.125−31.299
Intra-particle diffusion
Kp0.7630.7020.673
C15.27529.20536.997
R20.9790.8970.842
R2 adj0.9760.8820.820
MSE22.302131.873212.649
RMSE4.72311.48414.493
AIC29.68046.06749.975
Table 3. Isotherms studies for the dye adsorption onto chitosan and composites.
Table 3. Isotherms studies for the dye adsorption onto chitosan and composites.
Chitosan5 wt.% ZnO-Chitosan10 wt.% ZnO-Chitosan
Langmuir
Qm (mg g−1)68.07787.47190.976
KL (L mg−1)0.2220.4370.524
RL0.0430.2240.019
R20.9760.9820.983
R2 adj0.9690.9770.977
MSE0.0050.0020.001
RMSE0.0750.0450.039
AIC−24.881−29.558−31.091
Freundlich
KF (L g−1)13.42612.36812.075
nF4.1329.69917.241
R20.8680.7490.667
R2 adj0.8240.6610.556
MSE0.0140.0170.017
RMSE0.1170.1290.132
AIC−20.049−19.807−18.799
Temkin
bT (kJ mol−1)216.024205.905195.334
AT (L g−1)7.7455.0395.419
R20.9110.7950.718
R2 adj0.8820.7260.624
MSE43.84576.86388.748
RMSE6.6178.7679.421
AIC20.34923.15623.875
Table 4. Comparison of adsorption of various dyes onto chitosan and modified chitosan.
Table 4. Comparison of adsorption of various dyes onto chitosan and modified chitosan.
AdsorbentDyeConc (mg L−1) q e   ( mg   g 1 ) Equilibrium TimeRef
ChitosanReactive yellow
Reactive black
300-
-
800 min[42]
ChitosanMethylene blue10-4 h[43]
ChitosanMethylene blue109.8830 min[44]
ChitosanAcid dye
Basic dye
Direct dye
Reactive dye
10058.50
7.30
52.30
50.40
5 h[45]
Chitosan-cyclodextrinDirect blue 7830010.80350 min[46]
NCCA
NCCF
Red 601005.86
3.40
240 min[47]
Chitosan@Fe3O4Congo red10056.66600 min[48]
Chitosan/Al2O3/magnetiteMethyl orange2047.6040 min[49]
Sr3.8Fe25.7O70.4-chitosanCrystal violet
Basic red
5029.46
32.16
30 min[50]
ChitosanMethylene blue10088.77160 minThis work
5 wt.% ZnO-ChitosanMethylene blue193.78160 minThis work
10 wt.% ZnO-ChitosanMethylene blue10097.93160 minThis work
Chitosan-4-nitroacetophenone/CuO-CeO2-Al2O3 (NCCA); Chitosan-4-nitroacetophenon/CuO-CeO2-Fe2O3 (NCCF).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zango, Z.U.; Dennis, J.O.; Aljameel, A.I.; Usman, F.; Ali, M.K.M.; Abdulkadir, B.A.; Algessair, S.; Aldaghri, O.A.; Ibnaouf, K.H. Effective Removal of Methylene Blue from Simulated Wastewater Using ZnO-Chitosan Nanocomposites: Optimization, Kinetics, and Isotherm Studies. Molecules 2022, 27, 4746. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27154746

AMA Style

Zango ZU, Dennis JO, Aljameel AI, Usman F, Ali MKM, Abdulkadir BA, Algessair S, Aldaghri OA, Ibnaouf KH. Effective Removal of Methylene Blue from Simulated Wastewater Using ZnO-Chitosan Nanocomposites: Optimization, Kinetics, and Isotherm Studies. Molecules. 2022; 27(15):4746. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27154746

Chicago/Turabian Style

Zango, Zakariyya Uba, John Ojur Dennis, A. I. Aljameel, Fahad Usman, Mohammed Khalil Mohammed Ali, Bashir Abubakar Abdulkadir, Saja Algessair, Osamah A. Aldaghri, and Khalid Hassan Ibnaouf. 2022. "Effective Removal of Methylene Blue from Simulated Wastewater Using ZnO-Chitosan Nanocomposites: Optimization, Kinetics, and Isotherm Studies" Molecules 27, no. 15: 4746. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27154746

Article Metrics

Back to TopTop