Next Article in Journal
Alternative Controlling Agent of Theobroma grandiflorum Pests: Nanoscale Surface and Fractal Analysis of Gelatin/PCL Loaded Particles Containing Lippia origanoides Essential Oil
Previous Article in Journal
Fullerene Derivatives for Drug Delivery against COVID-19: A Molecular Dynamics Investigation of Dendro[60]fullerene as Nanocarrier of Molnupiravir
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chitosan-Functionalized Hydroxyapatite-Cerium Oxide Heterostructure: An Efficient Adsorbent for Dyes Removal and Antimicrobial Agent

1
Department of Chemistry, College of Science, Al-Baha University, P.O. Box 1988, Al-Baha 65799, Saudi Arabia
2
Department of Biology, College of Science, Al-Baha University, P.O. Box 1988, Al-Baha 65799, Saudi Arabia
3
Department of Chemistry, Faculty of Science and Arts, Al-Baha University, Qilwah 65941, Saudi Arabia
4
Lab. Desalination and Water Treatment Valorisation (LaDVEN), Water Research and Technologies Center (WRTC), BP 273, Soliman 8020, Tunisia
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(15), 2713; https://doi.org/10.3390/nano12152713
Submission received: 28 June 2022 / Revised: 29 July 2022 / Accepted: 2 August 2022 / Published: 7 August 2022
(This article belongs to the Section Biology and Medicines)

Abstract

:
The current research intended to employ a facile and economical process, which is also ecofriendly to transform camel waste bones into novel heterostructure for cleansing of diverse waste waters. The bones of camel were utilized for preparation of hydroxyapatite by hydrothermal method. The prepared hydroxyapatite was applied to the synthesis of cerium oxide-hydroxyapatite coated with natural polymer chitosan (CS-HAP-CeO2) heterostructure. Being abundant natural polymer polysaccharide, chitosan possesses exceptional assets such as accessibility, economic price, hydrophilicity, biocompatibility as well as biodegradability, therefore style it as an outstanding adsorbent for removing colorant and other waste molecules form water. This heterostructure was characterized by various physicochemical processes such as XRD, SEM-EDX, TEM, and FT-IR. The CS-HAP-CeO2 was screened for adsorption of various industrially important dyes, viz., Brilliant blue (BB), Congo red (CR), Crystal violet (CV), Methylene blue (MB), Methyl orange (MO), and Rhodamine B (RB) which are collective pollutants of industrial waste waters. The CS-HAP-CeO2 demonstrated exceptional adsorption against CR dye. The adsorption/or removal efficiency ranges are BB (11.22%), CR (96%), CV (28.22%), MB (47.74%), MO (2.43%), and RB (58.89%) dyes. Moreover, this heterostructure showed excellent bacteriostatic potential for E. coli, that is liable for serious waterborne diseases. Interestingly, this work revealed that the incorporation of cerium oxide and chitosan into hydroxyapatite substantially strengthened antimicrobial and adsorption capabilities than those observed in virgin hydroxyapatite. Herein, we recycled the unwanted camel bones into a novel heterostructure, which assists to reduce water pollution, mainly caused by the dye industries.

1. Introduction

The ecosphere encounters pure water crisis even in the 21st century. Unpolluted water is essential for the ecosystem; however, inattention and maladministration of natural water assets had considerably endangered the availability of pure water. Annually, water contamination and allied ailments prerogative affect approximately 2.1 million lives. Predominantly, various biological pollutants are accountable for water-related infections. The incidence of Coliforms is considered as indicator of fecal contagion. The eruptions of water-related microbial infections and inorganic materials are the main causes of loss of precious lives [1]. Admitting, not severely damaging in minute amounts, but they are the main pollutants once concentrated in water. Similarly, organic materials and prescription drugs etc., enter through different sources into water bodies [2]. Besides being aesthetic, the colored dyes are cancer-causing, and block light from entering aquatic systems. Numerous dyes pose a health risk since they are poisonous to both plants and animals [3]. Congo red (CR) is a benzidine, which is based on anionic disazo dyes. It is a human carcinogen. This dye is dangerous for both humans and the environment [4].
The waste-water treatment methods are very expensive. Bearing in mind, the probable claim for unsoiled water in Saudi Arabia, there is an imperative need for expansion of inexpensive materials for sanitization of contaminated waters. Therefore, current investigation visualizes the synthesis of economical material from Albaha camel bones. Camelus dromedarius, encompasses as dominant livestock of Arabian Peninsula. Due to appreciable consumption of camel meat in Saudi Arabia, huge amount of waste camel bones is being generated. These waste camel bones lead to hefty disposal costs and are major threat to health, aesthetic pleasure, and the environment. Therefore, in this study camel waste bones collected from Albaha region were utilized for the first time for purification of wastewater. Nevertheless, the chemical synthesis of hydroxyapatite (HAP) is complicated and expensive. Therefore, a very simple and low-cost procedure for the extraction of natural bone precursor HAP from waste camel bones is applied. The camel bone waste is a good natural source to develop hybrid nanocomposites which possess remarkable antimicrobial activity; hence, can be used to treat the household wastewaters such as kitchen, toilets, laundry etc. A plethora of studies have utilized animal bones for adsorption of mercury ions [5], removal of divalent heavy metals from wastewater [6], etc. Moreover, several organic/inorganic HAP composites are being utilized for removal of different dyes from waste waters. For instance, plain HAP, and composites of HAP have been used to adsorb CR dye [7]. Similarly, HAP doped with magnesium, sodium alginate was applied to remove Acid Yellow 220 [8]. Similarly, HAP-Chitosan composite is used for the elimination of CR, and HAP with polyalcohol is used to remove CR, MB, and MO [9]. It is reported in the literature that HAP is a noteworthy material for the elimination of toxic constituents from waste waters. Consequently, this study reports the synthesis of HAP from Albaha camel bones which were utilized for purification of wastewater particularly Saudi Arabia wastewater. The interesting properties of HAP [10] eases with opportunity to use it in numerous applications [11].
Similarly, chitosan (CS) can also be cast-off as an adsorbent to eliminate dyes owing to the occurrence of amino and hydroxyl clusters (which assist as energetic spots). CS is an abundant natural alkaline polymer polysaccharide. The exclusive chattels, for instance easiness, cheapness, water compatibility, biocompatibility, as well as biodegradability make it as an outstanding adsorbent for eliminating dyes [12,13,14,15,16]. The admirable functional groups of CS (amino and hydroxyl groups) craft it as a fabulous adsorbent for interaction with diverse dyes. Moreover, it possesses intrinsic biomedical applications such as CS is incorporated in dressings to reduce bleeding and as an antibacterial agent. CS can also be used to help to deliver drugs through skin [17,18]. Very recently, chitosan-Ulva lactuca composites for eliminating of Cd(II) ions from aquatic media have been employed which displayed outstanding efficacy [19]. Likewise, at present, the progressive semiconductor photocatalysis skill has been documented as an auspicious methodology to resolve water pollution and energy problems owing to its low-cost, eco-friendly, environment reusability and high removal efficiency characters. Semiconductor metal oxide nanophotocatalysts, such as CeO2, ZnO, SnO2, Bi2O3, and TiO2 have been widely used to eliminate the toxic pollutants in wastewater under different light sources [20]. In current times, cerium oxide nanoparticles have attracted increased attention owing to their shielded 4f electrons which is accountable for their captivating assets [21]. In an investigation, the synthesized CeO2 nanopowder, displayed extraordinary adsorption competence for elimination of diverse dyes at specific pH [22]. In particular, recently the CeO2 composites such as La2O3–CeO2–Fe3O4 nanofibers [23] and Ce/ZnO/CNFs catalysts exhibit good stability and reusability, which would be an economical and environmental friendly photocatalysts for various practical applications and can simply be recycled [20]. However, antagonistic characteristics viz., meager solubility of CeO2 can be overawed by packing CeO2 with biopolymer (such as chitosan, dextran, cyclodextrin, glucose and folic acid etc.,), therefore intensifies their miscibility and biocompatibility.
Consequently, with regard to exploring inexpensive material for purification of wastewater, this study recommends making usage of cheap and environment-friendly approaches to transform camel bone waste into novel composites such as CS-HAP-CeO2 heterostructure. The CS-HAP-CeO2 will then be applied as a chemical and biological agent for water purification. Moreover, solving water pollution crisis, our scientific contribution will also assist in plummeting the environmental pollution by recycling camel waste bones of Albaha province which might be a serious cause of pathogenic diseases in the very near future.

2. Materials and Methods

2.1. Preparation of Hydroxyapatite (HAP)

The camel bones were assembled from abattoir of Albaha city of Saudi Arabia. Extensive cleaning of bones was done following our earlier procedures [24], primarily, these mandibles were swept away extensively using water, also with acetone to exclude fat and superfluous impurities. Later on, bones were put in an oven at 170 °C for 48 h. Furthermore, it was calcined in a furnace at an elevated temperature of 700 °C to get HAP.

2.2. Synthesis of Chitosan Coated Cerium Oxide-HAP Heterostructure (CS-HAP-CeO2)

Cerium sulfate (5.6 g, anhydrous, 97%, Alfa Aesar, India) was added to 80 mL of water in a beaker. Then s sodium hydroxide solution (2 M) was added dropwise to attain pH~11. HAP (2 g) powder was mixed in the same solution with continuous stirring. About 10 mL of chitosan solution (500 mg of chitosan was liquefied in acetic acid solution, 1% v/v in water) was dissolved to it. At that time the final solution was shifted to the hydrothermal reactor and heated at 180 °C for 15 h. The prepared CS-HAP-CeO2 heterostructures were filtered and rinsed quite a few times with water and then with ethanol and kept at 90 °C overnight in an oven.

2.3. Characterization

The crystallinity of HAP and CS-HAP-CeO2 heterostructures were examined by X-ray diffraction (XRD, Rigaku (Tokyo, Japan) D/Max-2550, λ = 0.154 18 nm). The morphology of the samples was depicted by scanning electron microscopy equipped with energy-dispersive X-ray (SEM-EDX, JEOL, Tokyo, Japan). The presence of samples functional group was performed by Fourier transform infrared (FTIR, Thermo Scientific, Waltham, MA, USA) spectrometer. The microscopic images of samples were obtained by transmission electron microscopy (TEM, Hitachi, Tokyo, Japan).

2.4. Adsorption Studies

The adsorption study was conducted by batch method. The impact of various factors such as adsorbent dosage, contact time, pH, dyes concentration, and temperature on the removal efficiency of CR by CS-HAP-CeO2 heterostructure in the range (0.0025–0.1 g), (5–1440 min), (3–11 pH), (50–500 mg/L), and (25–45 °C) respectively, was tested. For adsorptive dye removal, 0.02 g of CS-HAP-CeO2 heterostructure adsorbent solution was thoroughly mixed with 50 mL aqueous dye solution of 50 mg/L in 100 mL conical flask. Dye pH was changed by the addition of 0.1 M HCl or 0.1 M NaOH and then placed on a shaker at 100 rpm at ambient temperature, and the samples were tested at specific time intervals. After equilibrium period, the solution was filtered and the concentration of CR, RB, MB, CV, BB, and MO dyes were observed by UV–visible spectrophotometer at maximum absorption wavelengths (λmax) of 500, 603, 665, 590, 592, and 468 nm, respectively. The removal efficiency (%) and adsorption capacity (mg/g) were determined using the following Equations (1) and (2), respectively [25].
Removal   efficiency   ( % ) = ( C o C e ) C o × 100
Adsorption   capacity   ( q e ) = ( C o C e ) × V m  
where Co is the initial concentration of dye and Ce is the concentration of MB equilibrium time), V is the volume of dye solutions, L, and m is the adsorbent dosage, g.

2.5. Antimicrobial Activity

The antibacterial activity of pristine HAP and novel CS-HAP-CeO2 heterostructure was verified in liquid broth [26]. The E. coli culture was maintained (106 CFU) with known amount of pristine HAP and CS-HAP-CeO2 heterostructure (0, 50, 100, and 200 µg/mL) to regulate the minimum inhibitory concentration (MIC). The expansion kinetics was supervised at every 4 h by reading the OD in a spectrophotometer. A persistent incubation temperature of 37 °C and rpm of 150 were sustained in a rotary shaker. Growth inhibition in presence of HAP and HAP-Ce coated with natural polymer CS was upheld to 16 h and the shift in absorbance was considered at 600 nm by ultra-violet [27] spectrophotometer.

3. Results

3.1. Sythesis and Characterization

The formation of CS-HAP-CeO2 heterostructure can be hypothesized as follows: the bone does have 65–70% HAP and 30–35% organic compounds (on dry weight basis). In our work, we extracted HAP by simple thermal treatment of natural bone. Cerium oxide was prepared using cerium sulfate precursors with HAP in solution. The use of sodium hydroxide resulted in a white precipitate, Ce(OH)3, from cerium sulfate solution. Oxidation of Ce3+ to Ce4+ in solution occurs at high pH and forms Ce(OH)4 [28]. Then it is converted to cerium oxide at high temperature. Furthermore, chitosan as a linear polysaccharide consists of (1,4)-linked 2-amino-deoxy-b-d-glucan with two free hydroxyl and one primary amino group. The positive and negative charge of amino and hydroxyl groups could make it available to attached with the surface of HAP and cerium oxide, thus resulting in the formation of novel CS-HAP-CeO2 heterostructure [28].
Figure 1 displays the XRD spectra of untainted HAP and CS-HAP-CeO2 heterostructure. All diffraction peaks of pure HAP correspond to single phase hexagonal structure of [Ca10 (PO4)6(OH)2] (JCPDS no. 09-0432). There was no contamination peaks present in the spectrum, it confirms the formation of pure HAP [29] (Figure 1a). The XRD patterns of CS-HAP-CeO2 heterostructure exhibited all characteristic diffraction peaks of HAP, in addition, it also shows the diffraction peaks corresponding to the crystalline plane (111), (200), (220), and (311) which confirms the cubic fluorite assembly of CeO2 (JCPDS no. 81-0792) [30]. In the same spectrum, a weak peak at around 2θ of 20° corresponds to pure chitosan [31] (Figure 1b).
Figure 2 presents the surface structure and morphology of pure HAP and CS-HAP-CeO2 heterostructure. Figure 2 illustrates the micro-structural study of HAP particles, which shows somewhat spherical in shape, displaying small and large particles with maximum size of around 1 µm (Figure 2a,b). Whereas, in the heterostructure, the HAP particles together with nanotube-like structure having diameter of around 100 nm with varying lengths (Figure 2c,d) were visible.
The close inspection of the TEM image confirmed the presence of CeO2 nanotubes having diameter of around 100 nm over the surface of HAP particles, which is in good agreement with SEM results (Figure 3a). The high-resolution image clearly shows different crystal structures on the heterostructure surface, the fringes of d = 0.34 nm and d = 0.27 nm, observed in Figure 3b matched well to those of (002) and (200) crystallographic planes of HAP and CeO2 particles, respectively. A thin amorphous uniform coating layer on the crystal surface can also be seen confirming the chitosan presence in the heterostructure (Figure 3b). This confirms that CS-HAP-CeO2 heterostructure was successfully prepared.
The EDX spectrum of CS-HAP-CeO2 showed the presence of all the expected elements in heterostructure. The EDX analysis confirms the presence of calcium (Ca), phosphorus (P), cerium (Ce), and oxygen (O) (Figure 4).

3.2. Adsorption Studies

3.2.1. Adsorbate Selectivity

The obtained adsorbent was tested for removal of various anionic dyes such as CR, RB, BB, and MO and cationic dyes such as CV and MB in aqueous medium at the parameters (dye = 20 mg/L, agitation = 100 rpm, dose = 0.02 g; and 25 °C) as shown in Figure 5. The experimental observations revealed that removal efficiency of the different dyes were BB (11.22%), CR (96%), CV (28.22%), MB (47.74%), MO (2.43%), and RB (58.89%). This indicates that the removal efficiency of CR dye was higher than the other tested dyes and order of affinity built on the amount of dye uptake was found to be: CR > RB > MB > CV > Bb > MO dye. This is owing to the fact that different dyes will undergo different physical and electrostatic forces, thus the difference in removal efficiency of dyes is due to their structure, molecular size, and functional groups which react with the functional groups of adsorbent surfaces with different physical and electrostatic forces [32,33]. Consequently, CR molecules occupy larger area over the CS-HAP-CeO2 heterostructure surface.

3.2.2. Effect of pH

The initial pH solution shows an important role in the adsorption process owing to its impact on both, the active sites of ionization process of the CR molecule and the CS-HAP-CeO2 heterostructure in the solution. Therefore, the effect of pH solution on the adsorption process at (CR = 100 mg/L, agitation = 100 rpm, dose = 0.02 g; and 25 °C) was tested at the varying initial pH (3–11) as presented in Figure 6a. The CR is a dipolar molecule and it represents a cationic form at acidic pH and an anionic form at basic pH [34]. Moreover, the normal red color of CR at pH around 7 changes to dark blue in a highly acidic medium and to red in basic medium ~pH (11–12), this color in an alkaline medium is slightly different from the original red at pH = 7 [35]. The results reveal that the removal efficiency and adsorption capacity increased from (242.84 mg/g and 97.13%) to (248.40 mg/g and 99.63%) with the increase in pH from 3 to 7, respectively, and then reduced. The dissociation constant (pKa) of CR is 4.0 [36]. According to this value when the pH < (pKa = 4), the surfaces of both CS-HAP-CeO2 heterostructure and CR dye are positively charged, showing an electrostatic repulsion between them. On the other hand, when the pH > (pKa = 4), the CR dissociates into polar groups (SO3) along with the positive charge of –N=N+– groups. The SO3 group over the CR surface adsorbs on positively charged CS-HAP-CeO2 adsorbent by electrostatic attraction along with H-bonding. As pH increases to 7, the removal efficiency of CR was increased to 99.84% owing to the number of positive charges in CS-HAP-CeO2 heterostructure surface, decreasing which led to creation of electrostatic attraction between the –N=N+– group of CR molecules and CS-HAP-CeO2 heterostructure surface. At an alkaline medium, at high pH, the existence of excess OH ions compete with the CR anions (sulfonate groups) for the adsorption sites on the CS-HAP-CeO2 heterostructure surface, leading to a decrease in the removal efficiency of CR dye [37,38]. Therefore, at pH = 7, probably, electrostatic attraction and H-bonding become dominant in the adsorption mechanism. Thus, pH 7 was chosen as the optimum pH value in further experiments. The reported effect of pH on adsorption of CR dye by MgO composite [39] is in agreement with our study. Furthermore, a recent review report based on the applications of artificial intelligence has captured incredible attention owing to solve future water-related problems [40]. Moreover, the cellulose nanofibril/rectorite composite sponges has displaced an efficient dye adsorption and selective separation in recent investigation [41].

3.2.3. Effect of Adsorbent Dose

The influence of CS-HAP-CeO2 heterostructure amount on CR dye adsorption at parameters conditions (CR = 20 mg/L, pH = 7; and 25 °C) in the range (0.0025–0.1 g) was studied as mentioned in Figure 6b. It was noticed that the removal efficiency of CR increases from 60.16% to 99.43% with the increase in amount of CS-HAP-CeO2 heterostructure adsorbent from 0.0025 to 0.02, and then became constant as the amount of CS-HAP-CeO2 heterostructure adsorbent increased to 0.1 g. An increase in the removal efficiency with the increase in amount of CS-HAP-CeO2 heterostructure adsorbent can be ascribed to the increasing accessibility of more adsorption sites [42]. On the other hand, the adsorption capacity was reduced from 240 mg/g to 20 mg/g with an increase in an adsorbent amount from 0.0025 to 0.1 g. This might be due to the aggregation events at high dosage [43]. Similar results were reported to the effect of adsorbent dose on adsorption of CR by Teucrium polium L. [44] and polypyrrole-modified red mud [45]. Therefore, 0.02 g was chosen as the optimal adsorbent dose under the examined conditions.

3.2.4. Effect of Contact Time

Figure 6c shows the effect of contact time on the removal of CR dye at an initial concentration of 100 mg/L onto CS-HAP-CeO2 adsorbent under various conditions (dose–0.02 g; pH–7; T–25 °C). The present outcome exhibited that the adsorption capacity of the CS-HAP-CeO2 heterostructure and removal efficiency of CR dye increased rapidly from (120.91 to 233.65 mg/g) and (48.36% to 96%) at 100 mg/L with the increase in time from 1 to 15 min at the initial concentration of CR dye, respectively. As a result, the ideal time for equilibrium for future work was evaluated to be 15 min. The CR adsorption on graphene–chitosan composite hydrogel has shown a comparable contact time effect [46].

3.2.5. Effect Temperatures

The effect of temperature on CR adsorption of dye on the CS-HAP-CeO2 adsorbent was investigated at different temperatures (25, 35, and 45 °C) and at different concentrations (50–400 mg/L). Figure 6d shows the initial CR dye concentration effect on the adsorption process at various concentrations (50 to 400 mg/L) and temperatures (25, 35, and 45 °C). The conditions are as follows: (dose–0.02 g; pH–7). When the CR dye concentration increased from 50 to 250 mg/L at 25 °C, the adsorption capacity of the CS-HAP-CeO2 adsorbent increased from 121.87 to 289.21 mg/g. The increase in the driving force for mass transfer from solution to the CS-HAP-CeO2 heterostructure adsorbent surface results in prominent increase in adsorption capacity at higher initial concentrations [47,48,49]. It was perceived that when the temperature rises, the adsorption capacity decreased marginally, indicating that the CR dye adsorption was an exothermic process. At higher temperatures, the binding forces on the surface with the CS-HAP-CeO2 heterostructure adsorbent were destroyed [50]. These results were consistent to that of CR adsorption on a fly ash/CeO2 composite adsorbent [51], malachite green and CR were removed from aqueous solution using magnetic HAP nanopowder [52], CR adsorption on graphene oxide/chitosan fibers [53], as well as CR adsorption on calcium HAP nanoparticles [54].

3.3. Adsorption Kinetics

To determine the CR adsorption mechanism onto CS-HAP-CeO2 heterostructure, two most common adsorption kinetics namely, pseudo-first-order equation (PFO) (Equation (3)) and pseudo-second-order rate equation (PSO) (Equation (4)) were applied. The equations can be expressed as follows [55]:
log ( q e q t ) = log ( q e ) k 1 2.303 t
t q t = 1 k 2 q e 2 + 1 q e t
where k1 (1/min) and k2 (g/(mg·min)) represent the PFO and PSO constant, respectively; qe and qt (mg/g) represent the adsorption capacity of the CS-HAP-CeO2 at equilibrium and at various times, respectively. The linear curves plotted of log (qeqt) and (t/qt) versus t are displayed in Figure 7, and the kinetic parameters are listed in Table 1.
The results show the experimental data fitted very well with PSO kinetic model, because the R2 value for the PFO (R2 = 0.9999) is higher than the R2 value for the PFO (R2 = 0.7967). Moreover, the adsorption capacity (qe,cal = 238.09 mg/g) value derived from PSO is close to that obtained experimentally (qe,exp = 237.95 mg/g). Hence, the results show that the adsorption of CR dye onto CS-HAP-CeO2 heterostructure is chemical adsorption via electrostatic attraction and H-bonding. The same result was observed for adsorption of CR by EG@MnFe2O4 and MnFe2O4 [36].

3.4. Adsorption Isotherm

To determine the mechanism of the adsorption process and evaluate the quantity of CR dye adsorption onto the CS-HAP-CeO2 heterostructure, four linear isotherm models namely, Langmuir Equation (5), Freundlich Equation (6), Temkin (Equations (7) and (8)), and Dubinin–Radushkevich (Equations (9)–(11)) were used. The equations can be expressed as follows [36]:
1 q e = 1 q m + 1 q m K L C e
log q e = log K f + 1 n log C e
q e = B ln A T + B ln C e
B = R T b T
ln q e = ln ( q s ) ( K a d ε 2 )
                          ε = R T ln ( 1 + 1 C e   )
E = 1 2 × K D
where Ce (mg/L) represents the equilibrium concentration of CR dye (mg/L), qm is the maximum monolayer adsorption capacity (mg/g), Kad (mol2/kJ2), Kf (mg/g), and KL (L/mg), represent the Dubinin–Radushkevich, Freundlich, and Langmuir isotherm constant, respectively; n is the adsorption intensity; AT (L/g) and B (J/mol) are the binding constant and the constant related to heat of sorption. ε is the Polanyi potential and E (kJ/mol) is the average free energy.
The linear plots of four isotherm models are shown in Figure 8, and the isotherm parameters are listed in Table 2. It was noticed that the R2 values of the Langmuir adsorption model were the highest (R2 = 0.9989) compared to the other isotherm models, Dubunin–Radushkevich (R2 = 0.9332), Temkin (R2 = 0.6384), and Freundlich (R2 = 0.5963). That means the adsorption process was well fitted with Langmuir adsorption model, representing monolayer adsorption on homogeneous surfaces. The maximum monolayer adsorption capacity of CS-HAP-CeO2 nanocomposite toward CR dye was 270.27 mg/g. This value is greater than the adsorption capacities of other adsorbents listed in Table 3 [34,56,57,58,59,60,61]. From the Dubinin–Radushkevich model, the magnitude of E is used for evaluating the type of adsorption mechanism. If the value of E is between 8 and 16 kJ/mol, it indicates chemisorption, while for values of E < 8 kJ/mol, the sorption process is of a physical nature [62,63]. Our results showed that the adsorption of CR on heterostructure adsorbent is physical adsorption, owing to the values of E being between 0.2672 and 1.118 kJ/mol (25 °C < T < 45 °C).

3.5. Thermodynamic Studies

To explain the effect of temperature on the CR adsorption over adsorbent and interpret whether the adsorption process occurs spontaneously or not, thermodynamic parameters including, free energy change (ΔG°), enthalpy change (ΔH°), and entropy change (ΔS°) were calculated using the following Equations (12) and (13) [64]:
Δ G ° = R T   ln K c
ln K c = Δ H ° R T + Δ S ° R
where k is Kc = qe/Ce. The values of ΔH° and ΔS° were calculated from the slope and intercept of the plot ln kc versus 1/T (Figure 9). The thermodynamic parameters are summarized in Table 4. The negative values of ΔG° and ΔH° imply that the CR adsorption over CS-HAP-CeO2 heterostructure are spontaneous and exothermic [65]. The values of ΔG° slightly increased from −11.63 to −5.69 (kJ mol−1) with the rising temperature from 298 to 318 K, indicating the adsorption process is favored at lower temperatures [65]. The ΔS° negative value implies a decrease in the randomness of the solid-solution interface.

3.6. CR Dye Adsorption Mechanism

CS-HAP-CeO2 heterostructure adsorbent contains functional groups on their surfaces such as phosphate, amino, and hydroxyl groups which are responsible for the binding of the CR dye. According to the dissociation constant (pKa) of CR dye, when the pH < (pKa = 4), the surfaces of both CS-HAP-CeO2 heterostructure and CR dye are positively charged, representing an electrostatic repulsion between them. On the other hand, when the pH > (pKa = 4), the CR dissociates into polar groups (SO3) along with the positive charge of the –N=N+– group. The SO3 group over CR surface adsorbs on positively charged CS-HAP-CeO2 adsorbent by electrostatic attraction along with H-bonding. As pH increases to 7, the removal efficiency of CR increased due to the decrease in the number of positive charges in the CS-HAP-CeO2 heterostructure surface which led to the creation of electrostatic attraction between the –N=N+– group of CR molecules and the negative charge of CS-HAP-CeO2 heterostructure surface. The functional groups of CS-HAP-CeO2 heterostructure before/and after adsorbing CR dye were identified through FTIR analysis, and the obtained results are shown in Figure 10. It was observed that the intensity of O–H/NH2 stretching shifted from 3529 to 3382 cm−1 and the band associated with P-O at 1038 cm−1 decreased in intensity and shifted to 1029 cm−1 after CR dye adsorption. This is due to the interaction of –N=N+– and –NH2, –SO3H groups on the CR surface of dye with PO4−3 and O–H/NH2 groups on the surface of CS-HAP-CeO2 heterostructure through electrostatic attraction and H- bonding. The appearance of two new bands at 830 and 755 cm−1 after CR dye adsorption indicates the CR dye loaded successfully onto the CS-HAP-CeO2 heterostructure surface. Therefore, the mechanism of adsorption of CR dye onto CS-HAP-CeO2 heterostructure occurred by two mechanisms namely, electrostatic interaction and H-bonding (Figure 11).

3.7. Antimicrobial

The virgin HAP and CS-HAP-CeO2 heterostructure were also evaluated using varying amounts against E. coli [66,67]. The produced heterostructure proved reputable antimicrobial activity against E. coli (MIC-50 μg/mL). The effects are shown in Figure 12. Alternatively, moderately trifling antibacterial achievement was detected with HAP at equal concentrations. Obviously, less bacteriostatic outcome was found at small amounts of HAP against E. coli, nonetheless improved activity was found with CS-HAP-CeO2 heterostructure. The greater activity of CS-HAP-CeO2 heterostructure was endorsed to morphological topographies with large surface [68] and synergism [69] between HAP and CeO2. The coating of CS also augmented in antimicrobial performance. The supposition about the antimicrobial action may be as: possibly at first HAP and CS-HAP-CeO2 heterostructure are interrelated with bacterial membrane, further disseminating into the inner cell prompting an outflow by disruption of the content of cell. Previous studies have described that nanocomposites via electrostatic attraction result in disruption of membrane, quash enzymes, break bacteria, ultimately hampering protein synthesis [70]. By and large, the bacteriostatic effect of HAP is correlated to liberation of OH ions in culture broth. Hydroxyl ions are tremendously oxidant free radicals which depict pronounced reactivity toward biomolecules. Furthermore, their lethal influence on bacterial cells are collectively allocated to aforesaid mechanisms [70].

4. Conclusions

In this paper, HAP and CS-HAP-CeO2 heterostructures were primed by hydrothermal method. The physicochemical data confirms efficacious formation of CS-HAP-CeO2 heterostructure. The adsorption results of CR dyes indicated outstanding efficiency of CS-HAP-CeO2 adsorbent. Additionally, the synthesized pristine and CS-HAP-CeO2 heterostructures unveiled good recyclable use and stability. The enhanced adsorption and antibacterial action of CS-HAP-CeO2 was mainly attributed to combined influence of CS, HAP and CeO2. Conclusively, outcomes of this study endorse an exceptionally capable adsorbent for decontamination of effluents; moreover, helps in reutilization of wastes of Baha city and meritoriously remove the environmental burden and govern the remediation procedure of the pathogenic diseases.

Author Contributions

Conceptualization, A.A.A., M.S.H., A.Q.A. and T.A.; methodology, A.A.A., A.Q.A. and M.S.H.; software, A.A.A. and T.A.; validation, A.A.A. and M.A.; formal analysis, A.A.A. and M.S.H.; investigation, A.A.A.; resources, A.A.A.; data curation, A.A.A. and M.A.; writing original draft preparation, A.A.A., M.S.H. and T.A.; supervision, A.A.A.; project administration, A.A.A.; funding acquisition, A.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are thankful to the Deanship of Scientific Research at Al-Baha University for funding this work under the General Research Funding program grant number 1442/15.

Institutional Review Board Statement

Not applicable.

Informed Con sent Statement

Not applicable.

Data Availability Statement

The data is included in the main text.

Acknowledgments

We gratefully acknowledge the encouragement and research support from the Deanship of Scientific research (Research grant number 1442/15) and the Dean Faculty of Science at Al-Baha University for encouragement and laboratory facilities throughout the research work.

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Helaly, M.N.; El-Metwally, M.A.; El-Hoseiny, H.; Omar, S.A.; El-Sheery, N.I. Effect of nanoparticles on biological contamination of ‘in vitro’cultures and organogenic regeneration of banana. Aust. J. Crop Sci. 2014, 8, 612. [Google Scholar]
  2. Nawab, J.; Khan, S.; Khan, M.A.; Sher, H.; Rehamn, U.U.; Ali, S.; Shah, S.M. Potentially toxic metals and biological contamination in drinking water sources in chromite mining-impacted areas of Pakistan: A comparative study. Expo. Health 2017, 9, 275–287. [Google Scholar] [CrossRef]
  3. Pathania, D.; Sharma, A.; Siddiqi, Z.-M. Removal of congo red dye from aqueous system using Phoenix dactylifera seeds. J. Mol. Liq. 2016, 219, 359–367. [Google Scholar] [CrossRef]
  4. Sathishkumar, K.; Alsalhi, M.S.; Sanganyado, E.; Devanesan, S.; Arulprakash, A.; Rajasekar, A. Sequential electrochemical oxidation and bio-treatment of the azo dye congo red and textile effluent. J. Photochem. Photobiol. B Biol. 2019, 200, 111655. [Google Scholar] [CrossRef]
  5. Dawlet, A.; Talip, D.; Mi, H.Y. Removal of mercury from aqueous solution using sheep bone charcoal. Procedia Environ. Sci. 2013, 18, 800–808. [Google Scholar] [CrossRef] [Green Version]
  6. Alqadami, A.A.; Khan, M.A.; Otero, M.; Siddiqui, M.R.; Jeon, B.-H.; Batoo, K.M. A magnetic nanocomposite produced from camel bones for an efficient adsorption of toxic metals from water. J. Clean. Prod. 2018, 178, 293–304. [Google Scholar] [CrossRef]
  7. Panneerselvam, K.; Arul, K.T.; Warrier, A.R.; Asokan, K.; Dong, C.-L. Rapid adsorption of industrial pollutants using metal ion doped hydroxyapatite. AIP Conf. Proc. 2019, 2117, 020004. [Google Scholar]
  8. Manatunga, D.C.; De Silva, R.M.; De Silva, K.N.; De Silva, N.; Premalal, E. Metal and polymer-mediated synthesis of porous crystalline hydroxyapatite nanocomposites for environmental remediation. R. Soc. Open Sci. 2018, 5, 171557. [Google Scholar] [CrossRef] [Green Version]
  9. Guan, Y.; Cao, W.; Guan, H.; Lei, X.; Wang, X.; Tu, Y.; Marchetti, A.; Kong, X. A novel polyalcohol-coated hydroxyapatite for the fast adsorption of organic dyes. Colloids Surf. A Physicochem. Eng. Asp. 2018, 548, 85–91. [Google Scholar] [CrossRef]
  10. Fihri, A.; Len, C.; Varma, R.S.; Solhy, A. Hydroxyapatite: A review of syntheses, structure and applications in heterogeneous catalysis. Coord. Chem. Rev. 2017, 347, 48–76. [Google Scholar] [CrossRef]
  11. Piccirillo, C.; Castro, P. Calcium hydroxyapatite-based photocatalysts for environment remediation: Characteristics, performances and future perspectives. J. Environ. Manag. 2017, 193, 79–91. [Google Scholar] [CrossRef] [PubMed]
  12. Tommalieh, M.; Ibrahium, H.A.; Awwad, N.S.; Menazea, A. Gold nanoparticles doped polyvinyl alcohol/chitosan blend via laser ablation for electrical conductivity enhancement. J. Mol. Struct. 2020, 1221, 128814. [Google Scholar] [CrossRef]
  13. Bagheri, A.R.; Ghaedi, M. Green preparation of dual-template chitosan-based magnetic water-compatible molecularly imprinted biopolymer. Carbohydr. Polym. 2020, 236, 116102. [Google Scholar] [CrossRef] [PubMed]
  14. Al-Kahtani, S.H.; Sofian, B.E.-D.E. Estimating preference change in meat demand in Saudi Arabia. Agric. Econ. 1995, 12, 91–98. [Google Scholar] [CrossRef]
  15. Gholami, H.; Ghaedi, M.; Arabi, M.; Ostovan, A.; Bagheri, A.R.; Mohamedian, H. Application of molecularly imprinted biomembrane for advancement of matrix solid-phase dispersion for clean enrichment of parabens from powder sunscreen samples: Optimization of chromatographic conditions and green approach. ACS Omega 2019, 4, 3839–3849. [Google Scholar] [CrossRef] [Green Version]
  16. Arabi, M.; Ostovan, A.; Bagheri, A.R.; Guo, X.; Wang, L.; Li, J.; Wang, X.; Li, B.; Chen, L. Strategies of molecular imprinting-based solid-phase extraction prior to chromatographic analysis. TrAC Trends Anal. Chem. 2020, 128, 115923. [Google Scholar] [CrossRef]
  17. Menazea, A.; Eid, M.; Ahmed, M. Synthesis, characterization, and evaluation of antimicrobial activity of novel Chitosan/Tigecycline composite. Int. J. Biol. Macromol. 2020, 147, 194–199. [Google Scholar] [CrossRef]
  18. Madni, A.; Kousar, R.; Naeem, N.; Wahid, F. Recent advancements in applications of chitosan-based biomaterials for skin tissue engineering. J. Bioresour. Bioprod. 2021, 6, 11–25. [Google Scholar] [CrossRef]
  19. Saad, E.M.; Elshaarawy, R.F.; Mahmoud, S.A.; El-Moselhy, K.M. New ulva lactuca algae based chitosan bio-composites for bioremediation of Cd (II) ions. J. Bioresour. Bioprod. 2021, 6, 223–242. [Google Scholar] [CrossRef]
  20. Jian, S.; Tian, Z.; Zhang, K.; Duan, G.; Yang, W.; Jiang, S. Hydrothermal synthesis of Ce-doped ZnO heterojunction supported on carbon nanofibers with high visible light photocatalytic activity. Chem. Res. Chin. Univ. 2021, 37, 565–570. [Google Scholar] [CrossRef]
  21. Al-Shwaiman, H.A.; Akshhayya, C.; Syed, A.; Bahkali, A.H.; Elgorban, A.M.; Das, A.; Varma, R.S.; Khan, S.S. Fabrication of intimately coupled CeO2/ZnFe2O4 nano-heterojunction for visible-light photocatalysis and bactericidal application. Mater. Chem. Phys. 2022, 279, 125759. [Google Scholar] [CrossRef]
  22. Tomic, N.M.; Dohcevic-Mitrovic, Z.D.; Paunović, N.M.; Mijin, D.A.Z.; Radić, N.D.; Grbic, B.V.; Askrabic, S.M.; Babić, B.M.; Bajuk-Bogdanovic, D.V. Nanocrystalline CeO2−δ as effective adsorbent of azo dyes. Langmuir 2014, 30, 11582–11590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Jian, S.; Shi, F.; Hu, R.; Liu, Y.; Chen, Y.; Jiang, W.; Yuan, X.; Hu, J.; Zhang, K.; Jiang, S. Electrospun magnetic La2O3–CeO2–Fe3O4 composite nanofibers for removal of fluoride from aqueous solution. Compos. Commun. 2022, 33, 101194. [Google Scholar] [CrossRef]
  24. Amna, T. Valorization of bone waste of Saudi Arabia by synthesizing hydroxyapatite. Appl. Biochem. Biotechnol. 2018, 186, 779–788. [Google Scholar] [CrossRef]
  25. Selim, S.E.; Meligi, G.A.; Abdelhamid, A.E.; Mabrouk, M.A.; Hussain, A.I. Novel Composite Films Based on Acrylic Fibers Waste/Nano-chitosan for Congo Red Adsorption. J. Polym. Environ. 2022, 30, 2642–2657. [Google Scholar] [CrossRef]
  26. Amna, T.; Hassan, M.S.; Barakat, N.A.; Pandeya, D.R.; Hong, S.T.; Khil, M.-S.; Kim, H.Y. Antibacterial activity and interaction mechanism of electrospun zinc-doped titania nanofibers. Appl. Microbiol. Biotechnol. 2012, 93, 743–751. [Google Scholar] [CrossRef] [PubMed]
  27. Guvensen, N.C.; Demir, S.; Ozdemir, G. Effects of magnesium and calcium cations on biofilm formation by Sphingomonas Paucimobilis from an industrial environment. Fresenius Environ. Bull. 2012, 21, 3685–3692. [Google Scholar] [CrossRef]
  28. Djuričić, B.; Pickering, S. Nanostructured cerium oxide: Preparation and properties of weakly-agglomerated powders. J. Eur. Ceram. Soc. 1999, 19, 1925–1934. [Google Scholar] [CrossRef]
  29. El Boujaady, H.; Mourabet, M.; El Rhilassi, A.; Bennani-Ziatni, M.; El Hamri, R.; Taitai, A. Adsorption of a textile dye on synthesized calcium deficient hydroxyapatite (CDHAp): Kinetic and thermodynamic studies. J. Mater. Environ. Sci. 2016, 7, 4049–4063. [Google Scholar]
  30. Nusrath, K.; Muraleedharan, K. Synthesis, characterization and thermal decomposition kinetics of cerium oxalate rods. Devagiri J. Sci. 2016, 2, 118–120. [Google Scholar]
  31. Kumar, S.; Koh, J. Physiochemical, optical and biological activity of chitosan-chromone derivative for biomedical applications. Int. J. Mol. Sci. 2012, 13, 6102–6116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Anirudhan, T.; Ramachandran, M. Adsorptive removal of basic dyes from aqueous solutions by surfactant modified bentonite clay (organoclay): Kinetic and competitive adsorption isotherm. Process Saf. Environ. Prot. 2015, 95, 215–225. [Google Scholar] [CrossRef]
  33. Allen, S.J.; Mckay, G.; Khader, K. Equilibrium adsorption isotherms for basic dyes onto lignite. J. Chem. Technol. Biotechnol. 1989, 45, 291–302. [Google Scholar] [CrossRef]
  34. Kaur, S.; Rani, S.; Mahajan, R.K. Adsorption Kinetics for the Removal of Hazardous Dye Congo Red by Biowaste Materials as Adsorbents. J. Chem. 2013, 2013, 628582. [Google Scholar] [CrossRef]
  35. Purkait, M.K.; Maiti, A.; Dasgupta, S.; De, S. Removal of congo red using activated carbon and its regeneration. J. Hazard. Mater. 2007, 145, 287–295. [Google Scholar] [CrossRef] [PubMed]
  36. Phạm, V.T.; Hong-Tham, N.; Tran, T.; Nguyen, D.; Le, H.; Nguyen, T.; Vo, D.-V.; Le, N.; Nguyen, D.C. Kinetics, Isotherm, Thermodynamics, and Recyclability of Exfoliated Graphene-Decorated MnFe2O4 Nanocomposite Towards Congo Red Dye. J. Chem. 2019, 2019, 5234585. [Google Scholar] [CrossRef]
  37. Astuti, D.; Aprilita, N.; Mudasir, M. Adsorption of the anionic dye of congo red from aqueous solution using a modified natural zeolite with benzalkonium chloride. Rasayan J. Chem. 2020, 13, 845–852. [Google Scholar] [CrossRef]
  38. Parvin, S.; Hussain, M.; Akter, F.; Biswas, B. Removal of Congo Red by Silver Carp (Hypophthalmichthys molitrix) Fish Bone Powder: Kinetics, Equilibrium, and Thermodynamic Study. J. Chem. 2021, 2021, 9535644. [Google Scholar] [CrossRef]
  39. Jiang, H.; Cao, Y.; Zeng, F.; Xie, Z.; He, F. A Novel Fe3O4/Graphene Oxide Composite Prepared by Click Chemistry for High-Efficiency Removal of Congo Red from Water. J. Nanomater. 2021, 2021, 9716897. [Google Scholar] [CrossRef]
  40. Alam, G.; Ihsanullah, I.; Naushad, M.; Sillanpää, M. Applications of artificial intelligence in water treatment for optimization and automation of adsorption processes: Recent advances and prospects. Chem. Eng. J. 2022, 427, 130011. [Google Scholar] [CrossRef]
  41. Chen, Y.; Hanshe, M.; Sun, Z.; Zhou, Y.; Mei, C.; Duan, G.; Zheng, J.; Shiju, E.; Jiang, S. Lightweight and anisotropic cellulose nanofibril/rectorite composite sponges for efficient dye adsorption and selective separation. Int. J. Biol. Macromol. 2022, 207, 130–139. [Google Scholar] [CrossRef] [PubMed]
  42. Çiner, F. Application of Fenton reagent and adsorption as advanced treatment processes for removal of Maxilon Red GRL. Glob. Nest J. 2018, 20, 1–6. [Google Scholar]
  43. Shoukat, S.; Bhatti, H.N.; Iqbal, M.; Noreen, S. Mango stone biocomposite preparation and application for crystal violet adsorption: A mechanistic study. Microporous Mesoporous Mater. 2017, 239, 180–189. [Google Scholar] [CrossRef]
  44. Alamrani, N.; Al_Aoh, H. Elimination of Congo Red Dye from Industrial Wastewater Using Teucrium polium L. as a Low-Cost Local Adsorbent. Adsorpt. Sci. Technol. 2021, 2021, 5728696. [Google Scholar] [CrossRef]
  45. Zhang, F.; Yin, Y.; Qiao, C.; Luan, Y.-N.; Guo, M.; Xiao, Y.; Liu, C. Anionic Dye Removal by Polypyrrole-Modified Red Mud and Its Application to a Lab-Scale Column: Adsorption Performance and Phytotoxicity Assessment. Adsorpt. Sci. Technol. 2021, 2021, 7694783. [Google Scholar] [CrossRef]
  46. Omidi, S.; Kakanejadifard, A. Eco-friendly synthesis of graphene–chitosan composite hydrogel as efficient adsorbent for Congo red. RSC Adv. 2018, 8, 12179–12189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Pirbazari, A.E.; Saberikhah, E.; Badrouh, M.; Emami, M.S. Alkali treated Foumanat tea waste as an efficient adsorbent for methylene blue adsorption from aqueous solution. Water Resour. Ind. 2014, 6, 64–80. [Google Scholar] [CrossRef] [Green Version]
  48. Nasuha, N.; Hameed, B.; Din, A.T.M. Rejected tea as a potential low-cost adsorbent for the removal of methylene blue. J. Hazard. Mater. 2010, 175, 126–132. [Google Scholar] [CrossRef]
  49. Al-Salihi, S.; Jasim, A.M.; Fidalgo, M.M.; Xing, Y. Removal of Congo red dyes from aqueous solutions by porous γ-alumina nanoshells. Chemosphere 2022, 286, 131769. [Google Scholar] [CrossRef]
  50. Al-Shehri, H.S.; Almudaifer, E.; Alorabi, A.Q.; Alanazi, H.S.; Alkorbi, A.S.; Alharthi, F.A. Effective adsorption of crystal violet from aqueous solutions with effective adsorbent: Equilibrium, mechanism studies and modeling analysis. Environ. Pollut. Bioavailab. 2021, 33, 214–226. [Google Scholar] [CrossRef]
  51. Ding, L.; Li, B.; Mi, J. The Effective Removal of Congo Red Dye from Aqueous Solution Using Fly Ash/CeO2 Composite Material. Appl. Mech. Mater. 2014, 535, 671–674. [Google Scholar] [CrossRef]
  52. Zhang, F.; Ma, B.; Jiang, X.; Ji, Y. Dual function magnetic hydroxyapatite nanopowder for removal of malachite green and Congo red from aqueous solution. Powder Technol. 2016, 302, 207–214. [Google Scholar] [CrossRef]
  53. Du, Q.; Sun, J.; Li, Y.; Yang, X.; Wang, X.; Wang, Z.; Xia, L. Highly enhanced adsorption of congo red onto graphene oxide/chitosan fibers by wet-chemical etching off silica nanoparticles. Chem. Eng. J. 2014, 245, 99–106. [Google Scholar] [CrossRef]
  54. Chahkandi, M. Mechanism of Congo red adsorption on new sol-gel-derived hydroxyapatite nano-particle. Mater. Chem. Phys. 2017, 202, 340–351. [Google Scholar] [CrossRef]
  55. Alqadami, A.A.; Naushad, M.; Abdalla, M.A.; Khan, M.R.; Alothman, Z.A. Adsorptive removal of toxic dye using Fe3O4–TSC nanocomposite: Equilibrium, kinetic, and thermodynamic studies. J. Chem. Eng. Data 2016, 61, 3806–3813. [Google Scholar] [CrossRef]
  56. Rao, R.; Jin, P.; Huang, Y.; Hu, C.; Dong, X.; Tang, Y.; Wang, F.; Luo, F.; Fang, S. A surface control strategy of CeO2 nanocrystals for enhancing adsorption removal of Congo red. Colloid Interface Sci. Commun. 2022, 49, 100631. [Google Scholar] [CrossRef]
  57. Azeez, L.; Adebisi, S.A.; Adejumo, A.L.; Busari, H.K.; Aremu, H.K.; Olabode, O.A.; Awolola, O. Adsorptive properties of rod-shaped silver nanoparticles-functionalized biogenic hydroxyapatite for remediating methylene blue and congo red. Inorg. Chem. Commun. 2022, 142, 109655. [Google Scholar] [CrossRef]
  58. Sirajudheen, P.; Karthikeyan, P.; Ramkumar, K.; Meenakshi, S. Effective removal of organic pollutants by adsorption onto chitosan supported graphene oxide-hydroxyapatite composite: A novel reusable adsorbent. J. Mol. Liq. 2020, 318, 114200. [Google Scholar] [CrossRef]
  59. Bensalah, H.; Younssi, S.A.; Ouammou, M.; Gurlo, A.; Bekheet, M.F. Azo dye adsorption on an industrial waste-transformed hydroxyapatite adsorbent: Kinetics, isotherms, mechanism and regeneration studies. J. Environ. Chem. Eng. 2020, 8, 103807. [Google Scholar] [CrossRef]
  60. Nguyen, N.T.; Nguyen, N.T.; Nguyen, V.A. In Situ Synthesis and Characterization of ZnO/Chitosan Nanocomposite as an Adsorbent for Removal of Congo Red from Aqueous Solution. Adv. Polym. Technol. 2020, 2020, 3892694. [Google Scholar] [CrossRef] [Green Version]
  61. Silva, V.C.; Araújo, M.E.B.; Rodrigues, A.M.; Vitorino, M.D.B.C.; Cartaxo, J.M.; Menezes, R.R.; Neves, G.A. Adsorption Behavior of Crystal Violet and Congo Red Dyes on Heat-Treated Brazilian Palygorskite: Kinetic, Isothermal and Thermodynamic Studies. Materials 2021, 14, 5688. [Google Scholar] [CrossRef] [PubMed]
  62. El Haddad, M. Removal of Basic Fuchsin dye from water using mussel shell biomass waste as an adsorbent: Equilibrium, kinetics, and thermodynamics. J. Taibah Univ. Sci. 2016, 10, 664–674. [Google Scholar] [CrossRef] [Green Version]
  63. Mladenovic, N.; Petkovska, J.; Dimova, V.; Dimitrovski, D.; Jordanov, I. Circular economy approach for rice husk modification: Equilibrium, kinetic, thermodynamic aspects and mechanism of Congo red adsorption. Cellulose 2022, 29, 503–525. [Google Scholar] [CrossRef]
  64. Cui, M.; Li, Y.; Sun, Y.; Wang, H.; Li, M.; Li, L.; Xu, W. Study on adsorption performance of MgO/Calcium alginate composite for congo red in wastewater. J. Polym. Environ. 2021, 29, 3977–3987. [Google Scholar] [CrossRef]
  65. Extross, A.; Waknis, A.; Tagad, C.; Gedam, V.; Pathak, P. Adsorption of congo red using carbon from leaves and stem of water hyacinth: Equilibrium, kinetics, thermodynamic studies. Int. J. Environ. Sci. Technol. 2022, in press. [Google Scholar] [CrossRef]
  66. Amna, T.; Hassan, M.S.; El-Newehy, M.H.; Alghamdi, T.; Moydeen Abdulhameed, M.; Khil, M.-S. Biocompatibility Computation of Muscle Cells on Polyhedral Oligomeric Silsesquioxane-Grafted Polyurethane Nanomatrix. Nanomaterials 2021, 11, 2966. [Google Scholar] [CrossRef]
  67. Amna, T.; Alghamdi, A.A.; Shang, K.; Hassan, M.S. Nigella Sativa-Coated Hydroxyapatite Scaffolds: Synergetic Cues to Stimulate Myoblasts Differentiation and Offset Infections. Tissue Eng. Regen. Med. 2021, 18, 787–795. [Google Scholar] [CrossRef]
  68. Ann, L.C.; Mahmud, S.; Bakhori, S.K.M.; Sirelkhatim, A.; Mohamad, D.; Hasan, H.; Seeni, A.; Rahman, R.A. Antibacterial responses of zinc oxide structures against Staphylococcus aureus, Pseudomonas aeruginosa and Streptococcus pyogenes. Ceram. Int. 2014, 40, 2993–3001. [Google Scholar] [CrossRef]
  69. Soletti, L.D.S.; Ferreira, M.E.C.; Kassada, A.T.; Abreu Filho, B.a.D.; Bergamasco, R.; Yamaguchi, N.U. Manganese ferrite graphene nanocomposite synthesis and the investigation of its antibacterial properties for water treatment purposes. Rev. Ambiente Água 2020, 15, e2515. [Google Scholar] [CrossRef]
  70. Goudarzi, M.R.; Bagherzadeh, M.; Fazilati, M.; Riahi, F.; Salavati, H.; Esfahani, S.S. Evaluation of antibacterial property of hydroxyapatite and zirconium oxide-modificated magnetic nanoparticles against Staphylococcus aureus and Escherichia coli. IET Nanobiotechnol. 2019, 13, 449–455. [Google Scholar] [CrossRef]
Figure 1. The XRD spectra of (a) pure HAP and (b) CS-HAP-CeO2 heterostructure. The symbol (●) indicates cerium oxide peaks whereas, (◆) represents chitosan peak.
Figure 1. The XRD spectra of (a) pure HAP and (b) CS-HAP-CeO2 heterostructure. The symbol (●) indicates cerium oxide peaks whereas, (◆) represents chitosan peak.
Nanomaterials 12 02713 g001
Figure 2. The SEM images of (a,b) pure HAP and (c,d) CS-HAP-CeO2 heterostructure. The arrows represent the CeO2 nanotubes embedded in HAP nanomatrix.
Figure 2. The SEM images of (a,b) pure HAP and (c,d) CS-HAP-CeO2 heterostructure. The arrows represent the CeO2 nanotubes embedded in HAP nanomatrix.
Nanomaterials 12 02713 g002
Figure 3. The (a) TEM (inset shows single nanotube) and (b) HR-TEM images of CS-HAP-CeO2 heterostructure.
Figure 3. The (a) TEM (inset shows single nanotube) and (b) HR-TEM images of CS-HAP-CeO2 heterostructure.
Nanomaterials 12 02713 g003
Figure 4. The EDX spectrum of CS-HAP-CeO2 heterostructure.
Figure 4. The EDX spectrum of CS-HAP-CeO2 heterostructure.
Nanomaterials 12 02713 g004
Figure 5. Selectivity studies of CS-HAP-CeO2 heterostructure toward different dyes.
Figure 5. Selectivity studies of CS-HAP-CeO2 heterostructure toward different dyes.
Nanomaterials 12 02713 g005
Figure 6. Effects of pH (a), adsorbent dosage (g) (b), contact time (min) (c), and initial concentrations (mg/L) (d) on the adsorption of CR dye onto CS-HAP-CeO2 heterostructure.
Figure 6. Effects of pH (a), adsorbent dosage (g) (b), contact time (min) (c), and initial concentrations (mg/L) (d) on the adsorption of CR dye onto CS-HAP-CeO2 heterostructure.
Nanomaterials 12 02713 g006
Figure 7. Plots of pseudo-first-order and pseudo-second-order kinetic model for adsorption of CR onto CS-HAP-CeO2 heterostructure adsorbents.
Figure 7. Plots of pseudo-first-order and pseudo-second-order kinetic model for adsorption of CR onto CS-HAP-CeO2 heterostructure adsorbents.
Nanomaterials 12 02713 g007
Figure 8. Linearized Irving Langmuir (a), Heurbet Freundlich (b), Mikhail Temkin (c), and Dubinin Radushkevich (d) isotherm adsorption models of CR dye onto CS-HAP-CeO2 heterostructure.
Figure 8. Linearized Irving Langmuir (a), Heurbet Freundlich (b), Mikhail Temkin (c), and Dubinin Radushkevich (d) isotherm adsorption models of CR dye onto CS-HAP-CeO2 heterostructure.
Nanomaterials 12 02713 g008
Figure 9. Plot of ln K vs. 1/T for thermodynamic parameters calculation.
Figure 9. Plot of ln K vs. 1/T for thermodynamic parameters calculation.
Nanomaterials 12 02713 g009
Figure 10. FTIR spectra of CS-HAP-CeO2 heterostructure before and after CR dye adsorption.
Figure 10. FTIR spectra of CS-HAP-CeO2 heterostructure before and after CR dye adsorption.
Nanomaterials 12 02713 g010
Figure 11. Mechanism of adsorption of CR onto CS-HAP-CeO2 heterostructure.
Figure 11. Mechanism of adsorption of CR onto CS-HAP-CeO2 heterostructure.
Nanomaterials 12 02713 g011
Figure 12. Bar illustrations demonstrating E. coli supplemented with different amounts of HAP and CS-HAP-CeO2 heterostructure (peak enlargement indicates E. coli deprived of HAP and CS-HAP-CeO2). Considerable difference (* p ≤ 0.05) was perceived between control and treatments. Substantial variance in bacteriostatic effect with HAP and CS-HAP-CeO2 at high quantity was envisaged. * p ≤ 0.05, *** p ≤ 0.001 suggestively different from virgin culture.
Figure 12. Bar illustrations demonstrating E. coli supplemented with different amounts of HAP and CS-HAP-CeO2 heterostructure (peak enlargement indicates E. coli deprived of HAP and CS-HAP-CeO2). Considerable difference (* p ≤ 0.05) was perceived between control and treatments. Substantial variance in bacteriostatic effect with HAP and CS-HAP-CeO2 at high quantity was envisaged. * p ≤ 0.05, *** p ≤ 0.001 suggestively different from virgin culture.
Nanomaterials 12 02713 g012
Table 1. Kinetic parameters for the pseudo-first and pseudo-second-order adsorption model.
Table 1. Kinetic parameters for the pseudo-first and pseudo-second-order adsorption model.
ModelParametersValue
Co: 100 mg/L, qe,exp.: 237.95 mg/g
Pseudo-first-orderqe1, cal. (mg/g)98.01
k1 (L/min)0.360
R20.7967
Pseudo-second-orderqe2, cal. (mg/g)238.09
k2 (g/mg-min)0.0098
R20.9999
Table 2. Parameters for plotting Langmuir, Freundlich, Temkin, and Dubinin–Radushkevich adsorption isotherms of CR onto CS-HAP-CeO2 heterostructure.
Table 2. Parameters for plotting Langmuir, Freundlich, Temkin, and Dubinin–Radushkevich adsorption isotherms of CR onto CS-HAP-CeO2 heterostructure.
ModelCR
298 K308 K318 K
Langmuir
qm, mg/g270.27238.09185.18
KL (L/mg)1.0000.5910.397
R20.99890.99870.9958
Freundlich
Kf,(mg/g) (L/mg)1/n158.19131.6784.58
N7.857.235.40
R20.59630.60270.6422
Dubinin-R
qs, mg/g282.56245.45196.56
KD-R (mol2 KJ−2)4.00 × 10−71.00 × 10−67.00 × 10−6
E (kJ mol−1)1.1180.70710.2672
R20.93320.98050.929
Temkin
bT = RT/B119.99130.15141.519
AT (L/g) 4029.702084.66340.63
B20.64719.03617.507
R20.63840.57250.4481
Table 3. Comparison of the maximum adsorption capacities for CR adsorption with different adsorbents.
Table 3. Comparison of the maximum adsorption capacities for CR adsorption with different adsorbents.
Adsorbentqm (mg/g)Isotherm/Kinetic ModelsRef.
CeO2 nanocrystals237Langmuir/PSO[56]
AgNPs-functionalized HAP159.11Freundlich/PSO[57]
CS@GO-Hap composite43.06Freundlich/PSO[58]
Hydroxyapatite (HAp)139 mg/gFreundlich/PSO[59]
ZnO/chitosan227.3Langmuir [60]
Ground nut shells charcoal (GNC)117.6Freundlich/PSO[34]
Eichhornia charcoal (EC)56.8Freundlich/PSO[34]
Polygorskite -700T136.1Freundlich/Elovich[61]
CS-HAP-CeO2270.27Langmuir/PSOThis study
Table 4. Thermodynamic parameters for the adsorption of CV on CS-HAP-CeO2 heterostructure adsorbent.
Table 4. Thermodynamic parameters for the adsorption of CV on CS-HAP-CeO2 heterostructure adsorbent.
Co (mg/L)(−) ΔH° (kJ/mol)(−) ΔS° (J/mol.K)(−) ΔG° (kJ/mol)
298 K308 K318 K
10019.438.2911.639.025.69
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alshahrani, A.A.; Alorabi, A.Q.; Hassan, M.S.; Amna, T.; Azizi, M. Chitosan-Functionalized Hydroxyapatite-Cerium Oxide Heterostructure: An Efficient Adsorbent for Dyes Removal and Antimicrobial Agent. Nanomaterials 2022, 12, 2713. https://doi.org/10.3390/nano12152713

AMA Style

Alshahrani AA, Alorabi AQ, Hassan MS, Amna T, Azizi M. Chitosan-Functionalized Hydroxyapatite-Cerium Oxide Heterostructure: An Efficient Adsorbent for Dyes Removal and Antimicrobial Agent. Nanomaterials. 2022; 12(15):2713. https://doi.org/10.3390/nano12152713

Chicago/Turabian Style

Alshahrani, Aisha A., Ali Q. Alorabi, M. Shamshi Hassan, Touseef Amna, and Mohamed Azizi. 2022. "Chitosan-Functionalized Hydroxyapatite-Cerium Oxide Heterostructure: An Efficient Adsorbent for Dyes Removal and Antimicrobial Agent" Nanomaterials 12, no. 15: 2713. https://doi.org/10.3390/nano12152713

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop