Next Article in Journal
Hyaluronan with Different Molecular Weights Can Affect the Gut Microbiota and Pathogenetic Progression of Post-Intensive Care Syndrome Mice in Different Ways
Next Article in Special Issue
Oxygen Nanocarriers for Improving Cardioplegic Solution Performance: Physico-Chemical Characterization
Previous Article in Journal
Effectiveness of Albumin-Fused Thioredoxin against 6-Hydroxydopamine-Induced Neurotoxicity In Vitro
Previous Article in Special Issue
The Osteogenic Properties of Calcium Phosphate Cement Doped with Synthetic Materials: A Structured Narrative Review of Preclinical Evidence
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of Box–Behnken Design to Optimize Phosphate Adsorption Conditions from Water onto Novel Adsorbent CS-ZL/ZrO/Fe3O4: Characterization, Equilibrium, Isotherm, Kinetic, and Desorption Studies

by
Endar Hidayat
1,2,
Nur Maisarah Binti Mohamad Sarbani
1,2,
Seiichiro Yonemura
1,2,
Yoshiharu Mitoma
1,2 and
Hiroyuki Harada
1,2,*
1
Graduate School of Comprehensive and Scientific Research, Prefectural University of Hiroshima, Shobara 727-0023, Japan
2
Department of Life and Environmental Science, Faculty of Bioresources Science, Prefectural University of Hiroshima, Shobara 727-0023, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(11), 9754; https://doi.org/10.3390/ijms24119754
Submission received: 11 May 2023 / Revised: 31 May 2023 / Accepted: 1 June 2023 / Published: 5 June 2023
(This article belongs to the Special Issue Multifunctional Application of Biopolymers and Biomaterials)

Abstract

:
Phosphate (PO43−) is an essential nutrient in agriculture; however, it is hazardous to the environment if discharged in excess as in wastewater discharge and runoff from agriculture. Moreover, the stability of chitosan under acidic conditions remains a concern. To address these problems, CS-ZL/ZrO/Fe3O4 was synthesized using a crosslinking method as a novel adsorbent for the removal of phosphate (PO43−) from water and to increase the stability of chitosan. The response surface methodology (RSM) with a Box–Behnken design (BBD)-based analysis of variance (ANOVA) was implemented. The ANOVA results clearly showed that the adsorption of PO43− onto CS-ZL/ZrO/Fe3O4 was significant (p ≤ 0.05), with good mechanical stability. pH, dosage, and time were the three most important factors for the removal of PO43−. Freundlich isotherm and pseudo-second-order kinetic models generated the best equivalents for PO43− adsorption. The presence of coexisting ions for PO43− removal was also studied. The results indicated no significant effect on PO43− removal (p ≤ 0.05). After adsorption, PO43− was easily released by 1 M NaOH, reaching 95.77% and exhibiting a good capability over three cycles. Thus, this concept is effective for increasing the stability of chitosan and is an alternative adsorbent for the removal of PO43− from water.

1. Introduction

Phosphate (PO43−) is a macronutrient needed for plant growth and is frequently applied as a fertilizer on agricultural lands. The increasing demands of food supply nowadays have led to the excessive application of fertilizer. However, excessive fertilizer use can cause PO43− to leach into waterways, leading to eutrophication and harmful algal bloom. These blooms diminish oxygen levels [1,2,3], interfere with aquatic life, and adversely affect the quality of drinking water (taste and odor) [4]. According to [5], PO43− decontamination must be performed efficiently while having a minimal impact on the surrounding ecosystem. Many methods have been reported to be effective in removing PO43− from water, including biological [6] methods, electrochemical [7,8] methods, precipitation [9], ion exchange [10], and adsorption [11,12]. Each strategy has advantages and disadvantages. Biological techniques are more economical; however, the residue of dead bacteria left behind after the process is inconvenient [13]. Electrochemical techniques are expensive but have a lower effectivity toward PO43− removal [14]. The precipitation process is simple and effective for chemical treatment but is inefficient for sewage sludge and waste disposal [15]. Ion exchange may also be used to remove anions by exchanging sulfates (SO42−) for PO43− ions; however, this would make the solution more corrosive, and it requires a costly clean-up (Blaney et al. [16]). Adsorption is the best option and is the most widely used method for water contaminants including PO43− ions [17,18]. This is because the technique is environmentally safe, the operation is easy and fast, and the technology is highly efficient.
Chitosan is currently gaining popularity as a potential adsorbent for water contaminants because it contains hydroxyl (–OH) and amino (–NH2) functional groups, which can easily react with other materials and are environmentally friendly [19]. This material, which cannot be accessed readily from nature, is synthesized through the chemical deacetylation of chitin. However, because of its low tensile strength and dissolution under acidic conditions, the use of chitosan directly in wastewater treatment technologies is not recommended. Therefore, chitosan must be modified to increase its chemical stability and adsorption capability [20]. The selection of an appropriate modification method and modifying agent is crucial for assessing the quality and functionality of the product created during the modification process. Crosslinking is one of the most frequently used procedures to enhance the physicochemical characteristics of chitosan [21,22]. Crosslinking is the process of combining two or more molecules via covalent bonds.
Zeolites are crystalline aluminum silicate (Al2O3·2SiO2) minerals with a porous and highly stable structure, and they could enhance the adsorption of chitosan onto their surface, leading to the improved stability of chitosan. These materials can be obtained from natural sources, such as shrimp, or can be synthesized using various methods [23]. Several reports have proven the use of chitosan and zeolite to remove dyes [24,25], pharmaceuticals [26], nitrate [27], and humic acid [28]. On the other hand, the fabrication of chitosan–metal oxides has attracted the attention of a lot of scientists owing to their numerous beneficial characteristics, such as chemical stability, a large surface area, and favorable adsorptive characteristics [29]. Magnesium oxide (MgO) [30], titanium oxide (TiO) [31], zinc oxide (ZnO) [32,33], zirconium oxide (ZrO) [34], and copper oxide (CuO) [35] are examples of metal oxides. ZrO was selected for this study owing to its strong affinity for anions [36].
The separation of the adsorbents is another issue of concern since the usual separation procedures result in the loss of the adsorbents as well as possible dangers to the environment [37,38]. Magnetite (Fe3O4) is one of the most magnetic particles that can be used in the manufacture of magnetic adsorbents for water purification because of its biodegradability, thermal stability, and large surface area [39,40]. The use of the crosslinking method to combine magnetite, zeolite, ZrO, and chitosan is a viable strategy. This is because the magnetic particles allow for easy separation when subjected to an external magnetic field, while the chitosan, zeolite, and ZrO provide many adsorption sites [41]. Therefore, the amalgamation of chitosan/zeolite/ZrO, and Fe3O4 (CS-ZL/ZrO/Fe3O4) may result in the development of novel composite materials with multifunctional constituents.
This study synthesized CS-ZL/ZrO/Fe3O4 with the target of using it as a novel adsorbent for PO43− removal from water. The response surface methodology (RSM) with the Box–Behnken design (BBD) optimization strategy was used to acquire insight into the effect of process factors such as pH, adsorbent dosage, temperature, and time to achieve the maximal adsorptive removal of PO43−. This process was performed to obtain the highest PO43− adsorptive removal. The adsorption isotherms and kinetic models were also calculated to figure out the adsorption mechanism.

2. Results and Discussion

2.1. Characterization of CS-ZL/ZrO/Fe3O4

The experimental results of BBD are listed in Table 1. It can be concluded that a pH of 2 offers the best conditions for PO43− removal. The pHZPC findings revealed that, at a pH of 2, the surface of CS-ZL/ZrO/Fe3O4 had a positive charge (pH < pHzpc) (Figure 1a). This might indicate the protonation of the -NH2 groups to -NH3+ groups on the surface. These attract negatively charged H2PO4 ions to CS-ZL/ZrO/Fe3O4, resulting in the construction of a surface complex between PO43− ions and CS-ZL/ZrO/Fe3O4. This study was similar to that reported by [42,43], which used SCBC-La and leftover coal, respectively, for PO43− removal under acidic conditions. The other possible reaction that could occur is shown in Equation (1).
Fe3O4 + 4Zr(OH)4 + 6H2PO4− → FeZr(PO4)3 + 12H2O
Table 2 summarizes the physical characteristics of these adsorbents. The results show that the BET-specific surface area was 88.1 m2/g, with a pore volume of 0.572 mL/g, an average diameter of 43.9 µm, and a porosity of 59%. These parameters show that the adsorbent had a substantial surface area for the adsorption of PO43− ions.
Figure 1b shows the XRD data used to verify the crystalline structure of the composite material. The XRD pattern shows a huge hump around 2θ = 21.22, which is a chitosan-specific peak [20]. Furthermore, the sharp peaks at 30.11, 35.49, 43.21 are mostly composed of crystalline phases, such as quartz, hematite, and alumina, which are all formed from zeolite- and zirconium-based materials. Magnetite corresponds to the peaks at 53.52, 57.08, and 62.78 [44]. Figure 1c shows a photograph of CS-ZL/ZrO/Fe3O4. It can be seen that the adsorbent is attached to the external magnet.
The SEM-EDS characterization of CS-ZL/ZrO/Fe3O4 was carried out to explore the surface properties and chemical components of the material. Figure 2 and Table 3 compare the SEM images and EDS data before and after PO43− adsorption. Before adsorption, the surface morphology of the adsorbent was sticky, rough, and porous. The surface became smooth and compact after PO43− adsorption, and this indicates that PO43− ions were trapped on the adsorbent surface. The primary objective of the EDS data analysis was to identify the components of the adsorbent materials. The weight percentages of Zr and Fe were the highest at 50.68 and 38.92%, respectively. The N value was derived from the chitosan materials [45,46,47]. Al, Si, and Fe were derived from zeolite and magnetite, respectively. Furthermore, the presence of P after the adsorption process indicates that PO43− was successfully adsorbed.
Figure 3 shows the functional groups in CS-ZL/ZrO/Fe3O4 before and after PO43− adsorption through an FTIR-ATR analysis. A CS-ZL/ZrO/Fe3O4 band was detected following PO43− adsorption from 3326 cm−1 to 3320 cm−1. This shows that PO43− ions interact with the stretching vibrations of hydrogen and amine in chitosan [48]. After PO43− adsorption, a decrease in the peak from 1634 to 1627 cm−1 was observed, which is associated with carboxyl groups (–COOCH3) [49]. An increased peak and a more curved and newer peak appeared after PO43− adsorption from 951 to 1006 cm−1 and at 2161 cm−1, which were assigned to Si-O-Al, Fe-O-Si, or Zr-O-Fe and CN stretching, respectively. This indicated a strong interaction with PO43− ions.

2.2. Mechanical Stability

The mechanical stability of the CS-ZL/ZrO/Fe3O4 composite was determined through the percentage of the initial mass that was preserved after drying. Figure 4a shows that increasing the concentration of the solution led to a higher WR%. Compared to the HCl-containing solution, the H2SO4-containing solution exhibited a higher WR%. Consequently, the crystalline structure of CS-ZL/ZrO/Fe3O4 was deformed, indicating that H2SO4 had significant contact with the chitosan group. Figure 4b shows the IR spectra after treatment. The positions of the peaks were consistent for all the samples. According to [50], the broad band visible at 3176–3345 cm−1 is assigned to the -NH2 groups changing to –NH3+ groups. The peaks between 1611 and 1630 cm−1, which were ascribed to the carboxyl (–COOCH3) and –NH2 groups, were generated through H+ generation by HCl and H2SO4. The peak shifted to 1068 cm−1, and expansion occurred when treated with 0.1 M H2SO4. SO42− ions have been shown to be associated with Si, Al, Fe, and Zr [51]. According to these results, the physical and chemical characteristics of the CS-ZL/ZrO/Fe3O4 composites did not change significantly.

2.3. ANOVA and Equations for Fitting Empirical Models

Table 4 shows the results of the statistical analysis, using the ANOVA test to evaluate the relationship between the input effective variables (A, B, C, and D) and their responses (Y). Table 4 shows that the F-value of the quadratic model was 16.68 and that the p-value was < 0.0001, indicating that this model was significant. Models A, B, D, A2, C2, D2, A × B, A × D, and C × D, marked with an asterisk (*), were found to be significant parameters of the model. The statistical variables obtained from the ANOVA test (Equation (2)) provide a full quadratic regression model for PO43− removal (%).
PO43− removal (%) = 99.2 − 1.72 A + 63 B − 1.478 C − 0.472 D + 0.2333 A2 + 840 B2 + 0.01575 C2 + 0.00661 D2 − 17.23 A*B − 0.0123A*C − 0.02107 A*D − 1.02 B*C + 2.098 B*D + 0.00343 C*D
The coefficients in the equation with positive and negative signs describe the additive and multiplicative effects of the variables on the response. The “Lack of Fit” was determined by comparing the residual error to the pure error in close proximity to the repeatedly designed points. F = 3.05 and p = 0.272 for the “Lack of Fit” revealed that it was not significant for the model (p < 0.05). It can be assumed that the model was adequately fitted and that there was no lack of fit.
The R2 value of the calculated second-order response model was 95.11, which is also known as the coefficient of determination. Consequently, it can be applied to reliably calculate the response at any given parameter level regardless of their values. In addition, a regression model was used to calculate the standardized influence of the independent factors on PO43− removal. A response surface plot was generated to investigate the influence of two components at initial the PO43− concentration of 20 mg/L (Figure 5). This plot was used to determine the standardized effects of all the independent variables. A surface plot is an easier way to display the response behavior that occurs when two parameters are simultaneously altered at the same time. It is more beneficial to select the quantity of various ingredients to obtain the desired response. Figure 5a displays a Pareto chart that compares the relative magnitude and the corresponding main, square, and interaction effects of the selected variables. The square effects of all four factors were found to be very significant (p ≤ 0.05) in addition to the main effects of the factors, which were also found to be highly significant (p ≤ 0.05). The ANOVA results reported in Table 4 led to the same conclusions. PO43− removal continuously increased in response to the pH, adsorbent dosage, and time. Figure 5b,c show that pH is a key factor in the removal of PO43−, and Figure 5d shows that increasing the contact time increases the percentage removal.

2.4. Initial Concentration and Isotherm Studies

The effects of the initial PO43− concentrations ranging from 20 to 500 mg/L, 0.06 g of adsorbent (CS-ZL/ZrO/Fe3O4), and pH (2.0) were investigated. Figure 6 shows that the PO43− adsorption capacity rose from 30.64 to 682.31 mg/g; however, the percentage of PO43− removal decreased from 91.91% to 81.88%. The adsorption capacity increased with the concentration because the total number of molecules increased. Consequently, the mass transfer resistance of adsorbate decreased. As a result, the percentage of removal decreased [52].
Adsorption isotherms are necessary to assess the capabilities of an adsorbent and the interactions between an adsorbate (a solution containing PO43− ions) and an adsorbent (CS-ZL/ZrO/Fe3O4). The acquired isotherm parameters can be used to conduct an accurate analysis while constructing an effective adsorption system. Both the equilibrium concentration and the adsorption capacity were estimated. The Langmuir model describes the monolayer adsorption processes at the designated homogenous surfaces on the adsorbent (Equation (3)). The essential property of the Langmuir model can be described as a dimensionless constant also known as the separation factor (RL), which is shown in Equation (4). By contrast, the Freundlich model is based on heterogeneous surfaces and multilayer sorption (Equation (5)). This is a linear equation, which is shown as follows:
C e / q e = ( C e q max ) + 1 / ( K 1 q max )
R L = ( 1 1 + bC o )  
Ln   q = lnK f + 1 n x   lnC e
qe (mg/g) is the adsorption capacity; K1 (L/mg) is the equilibrium constant of adsorption; qmax (mg/g) is the maximal adsorption capacity; Ce (mg/L) is the equilibrium concentration; Co (mg/L) is the initial concentration; RL is the separation factor; 0 < RL is favorable; RL > 1 is unfavorable; RL = 1 is linear; and Kf (mg/g) and 1/n are the adsorption capacity and the intensity of adsorption, respectively.
Figure 7 shows the isotherm model curves, and Table 5 shows the fitting results corresponding to these curves. The Freundlich model’s linear correlation coefficient (R2) was 0.9970, indicating that it provided the best fit compared to the other models. More importantly, the Langmuir and Freundlich parameters, known as RL and 1/n, indicate that the PO43− ion has a type of < 1. According to these data, the PO43− adsorption method is favorable.

2.5. Adsorption Kinetic Studies

This study investigated the influence of the contact time (35–2880 min) on PO43− adsorption at 30 °C. Figure 8 shows that the percentage of PO43− removal and the capacity for adsorption increased rapidly from 35 to 60 min and then gradually increased up to 90 min. This is because the adsorbent includes carboxyl, amine, hydrogen, and magnetite groups, all of which cause the adsorbent surface to become active and trap PO43− ions. Subsequently, the adsorption capacity decreased and increased, resulting in fast/slow PO43− adsorption, and it finally reached equilibrium at 1440 min, with an adsorption capacity and percent removal of 732.56 mg/g and 87.91%, respectively.
Adsorption kinetic studies are important because they deliver information on the adsorption mechanism, which is necessary to assess the effectiveness of the process [53]. Two kinetic models were used in this study: pseudo-first-order (PFO) (Equation (6)) and pseudo-second-order (PSO) (Equation (7)) models were investigated. The linear form can be obtained by calculating the following equation.
Log ( q e   q t ) = log q e K 1 t
t / q t = 1 / ( K 2 q e 2 ) + t / q e
where k1 (min−1) is the rate constant of the PFO model, t (min) is the time, and a linear plot of log t against log (qe − qt) and t against t/qt was used to determine K1 and K2 from the slope of the linear plots, respectively.
Figure 9 shows the fitting curves for the kinetic models, and Table 6 lists the fitting results corresponding to those curves. The findings confirm that the PSO model provided better results than the PFO model in terms of the linear correlation coefficient R2 value (0.9979). These findings imply that chemical processes control the adsorption rate.

2.6. Effect of Anions and Cations on PO43− Removal onto CS-ZL/ZrO/Fe3O4

Wastewater contains various substances, including anions and cations, which can affect the adsorption process [54]; it is essential to investigate the effect of ionic strength in an aqueous solution. This is because wastewater is made up of numerous components that might be found together. Figure 10 depicts the effect of different anions and cations on the PO43− adsorption capacity of CS-ZL/ZrO/Fe3O4. The experimental data indicate that there was no significant influence on PO43− removal. It revealed that the fabrication of CS-ZL/ZrO/Fe3O4 was effective in eliminating PO43− from water.

2.7. Desorption Studies

Figure 11a shows the desorption percentage of PO43− at different NaOH concentrations from 0.01 M to 1 M for 30 min at 30 °C. The results indicate that increasing the concentration increased the desorption percentage to 95.77%. Then, subsequent experiment at different contact times, from 30 to 150 min, using 1 M NaOH (Figure 11b). The desorption percentage increased and then decreased up to 150 min, which is similar to the results of the adsorption studies. The highest desorption percentage was observed after 30 min. The desorption mechanism may cause the hydroxide ions (OH-) in the sodium hydroxide solution to react with the CS-ZL/ZrO/Fe3O4-P surface and replace the PO43− groups, resulting in the release of PO43− into the liquid solution (Equation (8)). The reusability studies of PO43− adsorption onto CS-ZL/ZrO/Fe3O4 showed good performance for three cycles (Figure 11c).
H2PO4 + OH → HPO42− + H2O

2.8. Adsorption Performance Comparison

Table 7 compares the equilibrium and maximum adsorption capacity of CS-ZL/ZrO/Fe3O4 with those of various adsorbents. It can be seen that the pH is one of the main factors for PO43− removal onto the adsorbent, and the surface charge can become either positive or negative over a wide pH range, which influences the interaction between the adsorbent and PO43− ions. It is clear that the CS-ZL/ZrO/Fe3O4 adsorbent has a higher capacity than the other adsorbents. It is feasible to conclude that these adsorbents are viable alternatives for removing PO43− from water.

3. Materials and Methods

3.1. Materials

Chitosan (CH) (C6H11NO4) with molecular weight of 100,000–300,000 Da was bought from Acros Organics, Belgium. Zeolite (ZL) (Al2O3·2SiO2) was obtained from Tosoh Co. Ltd., Japan. Sodium hydroxide (NaOH), acetic acid (CH3COOH), disodium hydrogen phosphate (Na2HPO4), ferric chloride (FeCl3), ferrous sulfate (Fe2SO4), ammonium molybdate ((NH4)6Mo7O24·4H2O)), antimony potassium tartrate (K2Sb2(C4H2O6)2), ascorbic acid (C6H8O6), hydrochloric acid (HCl), and sulfuric acid (H2SO4) were bought from Kanto Chemical Co., Inc., Tokyo, Japan. ZrClO was purchased from Fujifilm Wako Chemical, Tokyo, Japan.

3.2. Synthesis of CS-ZL/ZrO/Fe3O4

CS-ZL/ZrO/Fe3O4 was synthesized through crosslinking method; chitosan (1 g) was dissolved in 100 mL of acetic acid (1%), and the resulting viscous solution was maintained at ambient temperature (25–30 °C) with magnetic stirring for 24 h (Equation (9)). Subsequently, 25 mL of the resulting chitosan solution was mixed with 0.5 g of zeolite and 20 mL of 1 M FeCl3 + 0.5 M Fe2SO4 + 0.5 M ZrClO. The mixture solution was then heated to 60 °C and was stirred for 1 h. The pH of the solution was adjusted to 10 using 3 M NaOH over 24 h with magnetic stirring at ambient temperature (25–30 °C), and the solution was filtered and washed multiple times with acetone and distilled water (DW) to remove any remaining NaOH. Subsequently, the materials were dried for 48 h in an oven at 60 °C (Equation (13)). The adsorbents are referred to as CS-ZL/ZrO/Fe3O4.
(CH3COOH)n + (C6H11NO4)m → (CH3COO)n(C6H11NO4H+)m
(C6H11NO4H+)n + Al2O3·2SiO2 → (C6H11NO4 − Al2O3·2SiO2)n + 2H2O
5FeCl3 + 15Fe2(SO4)3 + 12NaOH → 5Fe3O4 + 15Na2SO4 + 6H2O + 36NaCl
FeCl3 + 3Fe2(SO4)3 + ZrClO + 14NaOH → 5Fe3O4 + Zr(OH)4 + 2Na2SO4 + 6NaCl + 7H2O
2(CH3COO)n(C6H11NO4H+)m + 3Al2O3·2SiO2 + 3FeCl3 + Fe2SO4 + ZrClO4 + 14NaOH → [3Al2O3·2SiO2 −(C6H11NO4)]2m/3·Fe3O4·xH2O + 3Fe(OH)3 + 2Zr(OH)4 + 6NaCl + (2n + 2m)CH3COONa + (2n + m)H2O
Following this reaction, the negatively charged surface of the zeolite (Al2O3.2SiO2) may interact with the positively charged chitosan to produce chitosan–aluminosilicate complex. Electrostatic interactions between Fe3+ and Zr4+ ions and chitosan are another mechanism by which chitosan combines with metal ions to form chitosan–metal complexes. Fe(OH)3 and Fe3O4 are formed when Fe2+ and Fe3+ ions react with hydroxide ions (OH) from NaOH.

3.3. The Design of the Experiment

Experiments were conducted using response surface methodology (RSM) in combination with Box–Behnken design (BBD), and statistical analysis was performed using Minitab 21.3.1 software. (A) The pH (2–10), (B) dosage (0.02–0.1 g), (C) temperature (30–60 °C), and (D) contact time (10–60 min) were the independent variables examined in the BBD, with three levels and four parameters (Table 8). In total, 27 different sets of experiments were performed to determine the optimal conditions for PO43− removal. The data obtained were assessed using an equation for a quadratic polynomial response surface, which was calculated using Equation (14), to identify the relationships between independent variables and response.
Y = E 0 + E 1 A + E 2 B + E 3 C + E 4 D + E 11 A 2 + E 22 B 2   + E 33 C 2 + E 12 AB + E 13 AC + E 23 BC + ε
The coefficients of the polynomial model are represented as follows: E0 is constant expression, E1–E3 are linear effects, E11–E33 are second-order effects, E12–E23 are interactive effects, and ε is error term. An analysis of variance (ANOVA) was performed to calculate the F- and p-values of the model to measure its statistical significance and appropriateness. The statistical significance of the model is shown through the model’s F-value and p-value, and a lack-of-fit study of the proposed model was executed using Minitab 21.3.1 software. In addition, a 3D response surface plot and Pareto chart of standardized effects were developed to figure out the cooperative quantitative impact of the independent variables on the response and overall value of the model [63].

3.4. Batch Adsorption Study and Response Determination (PO43− Removal %)

To evaluate the efficiency of PO43− removal, batch adsorption approach was used in this study. In total, 100 mL of PO43− (20 mg/L) was placed in a 300 mL conical flask. After the adsorption procedure was completed, external magnetite was placed in the conical flask to separate the adsorbent and adsorbate. PO43− removal was calculated using Equation (15).
PO 4 3   removal   % = C o C e C o   ×   100
where Co and Ce are the initial and equilibrium PO43− concentrations (mg/L), respectively.
The data from run 17 of the BBD were used for subsequent experiments (isotherm and kinetic models). However, 30 min was not used because the results were far from equilibrium. The amount of PO43− adsorbed was determined using Equation (16).
q e = C o C e W   ×   V
where qe (mg/g) is the adsorption capacity, W (g) is the amount of CS-ZL/ZrO/Fe3O4, and V (L) is the volume of adsorbate (PO43− solution).

3.5. Adsorption Isotherm Studies

The isotherm model was studied with PO43− solutions ranging from 20 mg/L to 500 mg/L with pH of 2. These examinations were performed for 60 min at 30 °C, and adsorbent dosage of 0.06 g was placed in the flask. In this work, Langmuir and Freundlich models were used to assess PO43− adsorption onto CS-ZL/ZrO/Fe3O4 [64].

3.6. Adsorption Kinetic Studies

Pseudo-first-order (PFO) and pseudo-second-order (PSO) models were used to investigate the model of adsorption kinetics. The following parameters were used in the experiment: an adsorption temperature of 30 °C, an initial PO43− concentration of 500 mg/L at pH of 2, an adsorbent dosage of 0.06 g, and contact time ranging from 35 to 2880 min.

3.7. Influence of Coexisting Ionic Strength

The experiment was conducted under optimum conditions with a dosage of 0.06 g, an initial PO43− concentration of 500 mg/L, and a contact time of 1440 min at 30 °C. The coexisting ion was prepared with cationic and anionic ions at a concentration of 20 mg/L (Mg2+, Ca2+, CO32−, SO4, and Na+).

3.8. Desorption and Reusability Studies

In most practical applications, it is essential to employ adsorbents with high level of reusability. NaOH was chosen as desorbing agent to release PO43− ion from CS-ZL/ZrO/Fe3O4. Firstly, 0.06 g of CS-ZL/ZrO/Fe3O4 was loaded with 500 mg/L of PO43− ion at pH of 2.0, which was called CS-ZL/ZrO/Fe3O4-P. Then, 0.01 g of CS-ZL/ZrO/Fe3O4-P was dispersed in 60 mL of NaOH at 30 °C. The desorption capacity and desorption percentage are shown in Equations (17) and (18), respectively. Reusability was assessed using the same treatment as described above.
q des = C W   ×   V
%   Desorption = q des q e   ×   100
where qdes (mg/g) is the desorption capacity; C (mg/L) is the PO43− concentration of desorption; % Desorption (%) is the percentage desorption; and W, V, and qe are the same as above.

3.9. PO43− Measurements

PO43− ions were measured using the molybdate blue method. A total of 12 g of (NH4)6Mo7O24·4H2O was mixed with 100 mL of DW. K2Sb2(C4H2O6)2 (0.277 g) was added followed by 140 mL of 18 M H2SO4. Afterward, it was adjusted to 1 L with distilled water (solution A). A total of 1.06 g of C6H8O6 was added to and mixed with 100 mL of solution A, 25 mL of 4 N H2SO4 was added, and the solution was adjusted to 1 L with DW (solution B). Note: This solution must be prepared in every experiment. The procedure for the mixed solution was as follows: 2 mL of liquid sample/standard was mixed with 10 mL of solution B. Afterwards, we waited for 30 min and then analyzed the solution using a UV-Vis spectrophotometer (Jasco V-530) at a wavelength of 693 nm. A standard curve for PO43− was constructed using Na2HPO4.

3.10. Mechanical Stability

The mechanical stability of the CS-ZL/ZrO/Fe3O4 composite was evaluated based on the responses of the samples to a water bath shaker at 80 °C. For one hour, dried CS-ZL/ZrO/Fe3O4 was soaked in HCl and H2SO4 concentrations ranging from 0.01 to 0.1 M. Following that, the sample was dried in an oven at 60 °C for twenty-four hours. The calculation of the dry weight retention (WR) was performed using Equation (19).
WR   ( % ) = w i w a   ×   100
where wi and wa are the dry weights of CS-ZL/ZrO/Fe3O4 before and after treatment, respectively.

3.11. Characterization of CS-ZL/ZrO/Fe3O4

The crystalline structure of CS-ZL/ZrO/Fe3O4 was analyzed using a powder X-ray diffractometer (XRD) equipped with Cu/Kα radiation (Hypix-3000). Fourier transform infrared spectra (FTIR) of CS-ZL/ZrO/Fe3O4 were measured before and after PO43− adsorption using a Thermo Scientific Nicolet iS10 instrument (Thermo Fisher Scientific Inc., Waltham, MA, USA). The ATR-FTIR approach was used to analyze samples with a resolution of 4 cm−1 throughout the wavenumber spectrum spanning 400–4000 cm−1. To determine the specific surface area (SSA), the BET approach was combined with a surface area analyzer (MicroActive AutoPore V 9600 2.03.00, Micromeritics, Norcross, GA, USA). SEM-EDS (JIED-2300, Shimadzu, Kyoto, Japan) was used to examine the SEM images and the elemental distributions of CS-ZL/ZrO/Fe3O4. The initial (pHi) and final (pHf) pH values of the solutions were measured to determine the surface charge over a range of pH values (pHzpc). The pHi was adjusted from 2.0 to 10.0 in 0.01 M NaCl solution. Following that, 0.1 g of CS-ZL/ZrO/Fe3O4 was added and stirred for 24 h at 30 °C, and pHf was measured. A plot of ΔpH = pHf − pHi vs. pHi was used to determine pHpzc, which corresponds to the neutral surface charge.

3.12. Data Analysis

All results were noted and edited using Microsoft Excel. The effects of coexisting ions on PO43− removal were examined using a completely randomized design (CRD). Data were analyzed using ANOVA with Tukey’s test (p ≤ 0.05) using Minitab 21.3.1.

4. Conclusions

In this study, a novel adsorbent, CS-ZL/ZrO/Fe3O4, was prepared from chitosan (CS), zeolite (ZL), ZrO, and magnetite (Fe3O4) via a crosslinking approach. The Box–Behnken design (BBD) and the response surface methodology (RSM), with their corresponding four separate factors (pH, dosage, temperature, and time), were used to develop the best experimental conditions for PO43− removal. Weight retention (WR) was measured in a batch reactor under acidic conditions (HCl and H2SO4) at 80 °C for 1 h to determine the mechanical stability. The results indicate that CS-ZL/ZrO/Fe3O4 was stable and did not change in the functional group peak area after treatment. The best conditions were at a pH of 2.0, with an adsorption capacity and percentage removal of 732.56 mg/g and 87.91%, respectively. The Freundlich isotherm and pseudo-second-order (PSO) kinetic models were fitted to PO43− removal, indicating heterogeneous and chemical sorption. In addition, the results suggest that PO43− adsorption occurred via the electrostatic interactions between the positive charge of CS-ZL/ZrO/Fe3O4 and the negative charge of H2PO4− as well as ion exchange and hydrogen bonding. The presence of coexisting ions (Mg2+, Ca2+, CO32−, SO42−, and Na+) had no effect on the removal of PO43− (p ≤ 0.05). The desorption studies revealed that 1 M NaOH was better at releasing PO43−, reaching 95.77% after 30 min of treatment at 30 °C. The reusability of CS-ZL/ZrO/Fe3O4 showed good performance over three cycles. These findings imply that CS-ZL/ZrO/Fe3O4 is the best way to improve the stability of chitosan under acidic conditions, and it is a good adsorbent for removing PO43− and other potential water pollutants from water.

Author Contributions

Conceptualization, E.H.; Methodology, E.H.; Validation, Y.M.; Formal analysis, E.H.; Investigation, E.H. and Y.M.; Data curation, S.Y.; Writing—original draft, E.H.; Writing—review & editing, E.H. and N.M.B.M.S.; Visualization, H.H.; Supervision, S.Y., Y.M. and H.H.; Project administration, H.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

The author (E.H.) would like to express gratitude to the MEXT Scholarship for the funding received while studying at the Prefectural University of Hiroshima in Japan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yoon, S.-Y.; Lee, C.-G.; Park, J.-A.; Kim, J.-H.; Kim, S.-B.; Lee, S.-H.; Choi, J.-W. Kinetic, equilibrium and thermodynamic studies for phosphate adsorption to magnetic iron oxide nanoparticles. Chem. Eng. J. 2014, 236, 341–347. [Google Scholar] [CrossRef]
  2. Chen, M.; Ding, S.; Chen, X.; Sun, Q.; Fan, X.; Lin, J.; Ren, M.; Yang, L.; Zhang, C. Mechanisms driving phosphorus release during algal blooms based on hourly changes in iron and phosphorus concentrations in sediments. Water Res. 2018, 133, 153–164. [Google Scholar] [CrossRef] [PubMed]
  3. Schindler, D.W.; Carpenter, S.R.; Chapra, S.C.; Hecky, R.E.; Orihel, D.M. Reducing phosphorus to curb lake eutrophication is a success. Environ. Sci. Technol. 2016, 50, 8923–8929. [Google Scholar] [CrossRef] [Green Version]
  4. Pawar, R.R.; Gupta, P.; Lalhmunsiama; Bajaj, H.C.; Lee, S.-M. Al-intercalated acid activated bentonite beads for the removal of aqueous phosphate. Sci. Total Environ. 2016, 572, 1222–1230. [Google Scholar] [CrossRef]
  5. Xie, Q.; Li, Y.; Lv, Z.; Zhou, H.; Yang, X.; Chen, J.; Guo, H. Effective Adsorption and Removal of Phosphate from Aqueous Solutions and Eutrophic Water by Fe-based MOFs of MIL-101. Sci. Rep. 2017, 7, 3316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Wei, W.; Du, J.; Li, J.; Yan, M.; Zhu, Q.; Jin, X.; Zhu, X.; Hu, Z.; Tang, Y.; Lu, Y. Construction of Robust Enzyme Nanocapsules for Effective Organophosphate Decontamination, Detoxification, and Protection. Adv. Mater. 2013, 25, 2212–2218. [Google Scholar] [CrossRef]
  7. Wang, F.; Wei, J.; Zou, X.; Fu, R.; Li, J.; Wu, D.; Lv, H.; Zhu, G.; Wu, X.; Chen, H. Enhanced electrochemical phosphate recovery from livestock wastewater by adjusting pH with plant ash. J. Environ. Manag. 2019, 250, 109473. [Google Scholar] [CrossRef]
  8. Ren, Y.; Zheng, W.; Duan, X.; Goswami, N.; Liu, Y. Recent advances in electrochemical removal and recovery of phosphorus from water: A review. Environ. Funct. Mater. 2022, 1, 10–20. [Google Scholar] [CrossRef]
  9. Hidayat, E.; Harada, H. Simultaneously Recovery of Phosphorus and Potassium Using Bubble Column Reactor as Struvite-K and Implementation on Crop Growth. In Crystallization and Applications; IntechOpen: London, UK, 2022. [Google Scholar] [CrossRef]
  10. Guida, S.; Rubertelli, G.; Jefferson, B.; Soares, A. Demonstration of ion exchange technology for phosphorus removal and recovery from municipal wastewater. Chem. Eng. J. 2021, 420, 129913. [Google Scholar] [CrossRef]
  11. Hidayat, E.; Yoshino, T.; Yonemura, S.; Mitoma, Y.; Harada, H. A Carbonized Zeolite/Chitosan Composite as an Adsorbent for Copper (II) and Chromium (VI) Removal from Water. Materials 2023, 16, 2532. [Google Scholar] [CrossRef]
  12. Zahari, K.F.A.; Sahu, U.K.; Khadiran, T.; Surip, S.N.; Alothman, Z.A.; Jawad, A.H. Mesoporous Activated Carbon from Bamboo Waste via Microwave-Assisted K2CO3 Activation: Adsorption Optimization and Mechanism for Methylene Blue Dye. Separations 2022, 9, 390. [Google Scholar] [CrossRef]
  13. Matei, A.; Racoviteanu, G. Review of the technologies for nitrates removal from water intended for human consumption. In IOP Conference Series: Earth and Environmental Science; IOP Publishing Ltd.: Bristol, UK, 2021. [Google Scholar] [CrossRef]
  14. Rodrigues, L.A.; da Silva, M.L.C.P. An investigation of phosphate adsorption from aqueous solution onto hydrous niobium oxide prepared by co-precipitation method. Colloids Surf. A Physicochem. Eng. Asp. 2009, 334, 191–196. [Google Scholar] [CrossRef]
  15. Huang, W.; Zhang, Y.; Li, D. Adsorptive removal of phosphate from water using mesoporous materials: A review. In Journal of Environmental Management; Academic Press: Cambridge, MA, USA, 2017; Volume 193, pp. 470–482. [Google Scholar]
  16. Blaney, L.; Cinar, S.; Sengupta, A.K. Hybrid anion exchanger for trace phosphate removal from water and wastewater. Water Res. 2007, 41, 1603–1613. [Google Scholar] [CrossRef] [PubMed]
  17. Hidayat, E.; Yoshino, T.; Yonemura, S.; Mitoma, Y.; Harada, H. Synthesis, Adsorption Isotherm and Kinetic Study of Alkaline- Treated Zeolite/Chitosan/Fe3+ Composites for Nitrate Removal from Aqueous Solution—Anion and Dye Effects. Gels 2022, 8, 782. [Google Scholar] [CrossRef] [PubMed]
  18. Almasri, D.A.; Saleh, N.B.; Atieh, M.A.; McKay, G.; Ahzi, S. Adsorption of phosphate on iron oxide doped halloysite nanotubes. Sci. Rep. 2019, 9, 3232. [Google Scholar] [CrossRef] [PubMed]
  19. Ladeira, N.M.B.; Donnici, C.L.; de Mesquita, J.P.; Pereira, F.V. Preparation and characterization of hydrogels obtained from chitosan and carboxymethyl chitosan. J. Polym. Res. 2021, 28, 335. [Google Scholar] [CrossRef]
  20. Mohammed, I.A.; Malek, N.N.A.; Jawad, A.H.; Mastuli, M.S.; Alothman, Z.A. Box–Behnken Design for Optimizing Synthesis and Adsorption Conditions of Covalently Crosslinked Chitosan/Coal Fly Ash Composite for Reactive Red 120 Dye Removal. J. Polym. Environ. 2022, 30, 3447–3462. [Google Scholar] [CrossRef]
  21. Jawad, A.H.; Hameed, B.H.; Abdulhameed, A.S. Synthesis of biohybrid magnetic chitosan-polyvinyl alcohol/MgO nanocomposite blend for remazol brilliant blue R dye adsorption: Solo and collective parametric optimization. Polym. Bull. 2022, 80, 4927–4947. [Google Scholar] [CrossRef]
  22. Aramesh, N.; Bagheri, A.R.; Bilal, M. Chitosan-based hybrid materials for adsorptive removal of dyes and underlying interaction mechanisms. In International Journal of Biological Macromolecules; Elsevier B.V.: Amsterdam, The Netherlands, 2021; Volume 183, pp. 399–422. [Google Scholar] [CrossRef]
  23. Salehi, S.; Anbia, M. Adsorption Selectivity of CO2 and CH4 on Novel PANI/Alkali-Exchanged FAU Zeolite Nanocomposites. J. Inorg. Organomet. Polym. Mater. 2017, 27, 1281–1291. [Google Scholar] [CrossRef]
  24. Hidayat, E.; Yonemura, S.; Mitoma, Y.; Harada, H. Methylene Blue Removal by Chitosan Cross-Linked Zeolite from Aqueous Solution and Other Ion Effects: Isotherm, Kinetic, and Desorption Studies. Adsorpt. Sci. Technol. 2022, 2022, 1853758. [Google Scholar] [CrossRef]
  25. Hidayat, E.; Harada, H.; Mitoma, Y.; Yonemura, S.; A Halem, H.I. Rapid Removal of Acid Red 88 by Zeolite/Chitosan Hydrogel in Aqueous Solution. Polymers 2022, 14, 893. [Google Scholar] [CrossRef]
  26. Vakili, M.; Qiu, W.; Cagnetta, G.; Huang, J.; Yu, G. Mechanochemically oxidized chitosan-based adsorbents with outstanding Penicillin G adsorption capacity. J. Environ. Chem. Eng. 2021, 9, 105454. [Google Scholar] [CrossRef]
  27. Gao, Y.; Bao, S.; Zhang, L.; Zhang, L. Nitrate removal by using chitosan/zeolite molecular sieves composite at low temperature: Characterization, mechanism, and regeneration studies. Desalination Water Treat 2020, 203, 160–171. [Google Scholar] [CrossRef]
  28. Lin, J.; Zhan, Y. Adsorption of humic acid from aqueous solution onto unmodified and surfactant-modified chitosan/zeolite composites. Chem. Eng. J. 2012, 200–202, 202–213. [Google Scholar] [CrossRef]
  29. Mohammad, A.K.T.; Abdulhameed, A.S.; Jawad, A.H. Box-Behnken design to optimize the synthesis of new crosslinked chitosan-glyoxal/TiO2 nanocomposite: Methyl orange adsorption and mechanism studies. Int. J. Biol. Macromol. 2019, 129, 98–109. [Google Scholar] [CrossRef] [PubMed]
  30. Nga, N.K.; Chau, N.T.T.; Viet, P.H. Preparation and characterization of a chitosan/MgO composite for the effective removal of reactive blue 19 dye from aqueous solution. J. Sci. Adv. Mater. Devices 2020, 5, 65–72. [Google Scholar] [CrossRef]
  31. Spoială, A.; Ilie, C.-I.; Dolete, G.; Croitoru, A.-M.; Surdu, V.-A.; Trușcă, R.-D.; Motelica, L.; Oprea, O.-C.; Ficai, D.; Ficai, A.; et al. Preparation and Characterization of Chitosan/TiO2 Composite Membranes as Adsorbent Materials for Water Purification. Membranes 2022, 12, 804. [Google Scholar] [CrossRef]
  32. Reghioua, A.; Barkat, D.; Jawad, A.H.; Abdulhameed, A.S.; Rangabhashiyam, S.; Khan, M.R.; Alothman, Z.A. Magnetic Chitosan-Glutaraldehyde/Zinc Oxide/Fe3O4 Nanocomposite: Optimization and Adsorptive Mechanism of Remazol Brilliant Blue R Dye Removal. J. Polym. Environ. 2021, 29, 3932–3947. [Google Scholar] [CrossRef]
  33. Reghioua, A.; Barkat, D.; Jawad, A.H.; Abdulhameed, A.S.; Khan, M.R. Synthesis of Schiff’s base magnetic crosslinked chitosan-glyoxal/ZnO/Fe3O4 nanoparticles for enhanced adsorption of organic dye: Modeling and mechanism study. Sustain. Chem. Pharm. 2021, 20, 100379. [Google Scholar] [CrossRef]
  34. Cho, D.-W.; Jeon, B.-H.; Jeong, Y.; Nam, I.-H.; Choi, U.-K.; Kumar, R.; Song, H. Synthesis of hydrous zirconium oxide-impregnated chitosan beads and their application for removal of fluoride and lead. Appl. Surf. Sci. 2016, 372, 13–19. [Google Scholar] [CrossRef]
  35. Sathiyavimal, S.; Vasantharaj, S.; Kaliannan, T.; Pugazhendhi, A. Eco-biocompatibility of chitosan coated biosynthesized copper oxide nanocomposite for enhanced industrial (Azo) dye removal from aqueous solution and antibacterial properties. Carbohydr. Polym. 2020, 241, 116243. [Google Scholar] [CrossRef]
  36. Salehi, S.; Alijani, S.; Anbia, M. Enhanced adsorption properties of zirconium modified chitosan-zeolite nanocomposites for vanadium ion removal. Int. J. Biol. Macromol. 2020, 164, 105–120. [Google Scholar] [CrossRef] [PubMed]
  37. Rosales, E.; Anasie, D.; Pazos, M.; Lazar, I.; Sanromán, M.A. Kaolinite adsorption-regeneration system for dyestuff treatment by Fenton based processes. Sci. Total Environ. 2018, 622–623, 556–562. [Google Scholar] [CrossRef] [PubMed]
  38. Ghobadi, M.; Gharabaghi, M.; Abdollahi, H.; Boroumand, Z.; Moradian, M. MnFe2O4-graphene oxide magnetic nanoparticles as a high-performance adsorbent for rare earth elements: Synthesis, isotherms, kinetics, thermodynamics and desorption. J. Hazard. Mater. 2018, 351, 308–316. [Google Scholar] [CrossRef]
  39. Maksoud, M.I.A.A.; Elgarahy, A.M.; Farrell, C.; Al-Muhtaseb, A.H.; Rooney, D.W.; Osman, A.I. Insight on water remediation application using magnetic nanomaterials and biosorbents. In Coordination Chemistry Reviews; Elsevier B.V.: Amsterdam, The Netherlands, 2020; Volume 403. [Google Scholar]
  40. Fu, Z.; Li, H.; Yang, L.; Yuan, H.; Jiao, Z.; Chen, L.; Huang, J.; Liu, Y.-N. Magnetic polar post-cross-linked resin and its adsorption towards salicylic acid from aqueous solution. Chem. Eng. J. 2015, 273, 240–246. [Google Scholar] [CrossRef]
  41. Al-Harahsheh, M.; AlJarrah, M.; Mayyas, M.; Alrebaki, M. High-stability polyamine/amide-functionalized magnetic nanoparticles for enhanced extraction of uranium from aqueous solutions. J. Taiwan Inst. Chem. Eng. 2018, 86, 148–157. [Google Scholar] [CrossRef]
  42. Zong, E.; Shen, Y.; Yang, J.; Liu, X.; Song, P. Preparation and Characterization of an Invasive Plant-Derived Biochar-Supported Nano-Sized Lanthanum Composite and Its Application in Phosphate Capture from Aqueous Media. ACS Omega 2023, 8, 14177–14189. [Google Scholar] [CrossRef]
  43. Mekonnen, D.T.; Alemayehu, E.; Lennartz, B. Removal of Phosphate Ions from Aqueous Solutions by Adsorption onto Leftover Coal. Water 2020, 12, 1381. [Google Scholar] [CrossRef]
  44. Zhang, M.; Zhang, Z.; Peng, Y.; Feng, L.; Li, X.; Zhao, C.; Sarfaraz, K. Novel cationic polymer modified magnetic chitosan beads for efficient adsorption of heavy metals and dyes over a wide pH range. Int. J. Biol. Macromol. 2020, 156, 289–301. [Google Scholar] [CrossRef]
  45. Safaei-Ghomi, J.; Tavazo, M.; Shahbazi-Alavi, H. Chitosan-attached nano-Fe3O4 as a superior and retrievable heterogeneous catalyst for the synthesis of benzopyranophenazines using chitosan-attached nano-Fe3O4. Z. Fur Nat.—Sect. B J. Chem. Sci. 2019, 74, 733–738. [Google Scholar] [CrossRef]
  46. Annaduzzaman, M.; Bhattacharya, P.; Ersoz, M.; Lazarova, Z. Characterization of a chitosan biopolymer and arsenate removal for drinking water treatment. In One Century of the Discovery of Arsenicosis in Latin America (1914–2014): As 2014—Proceedings of the 5th International Congress on Arsenic in the Environment; CRC Press/Balkema: Boca Raton, FL, USA, 2014; pp. 745–747. [Google Scholar]
  47. Yang, T.; Zhang, W.; Liu, H.; Guo, Y. Enhanced removal of U(VI) from aqueous solution by chitosan-modified zeolite. J. Radioanal. Nucl. Chem. 2020, 323, 1003–1012. [Google Scholar] [CrossRef]
  48. Fajardo, A.R.; Lopes, L.C.; Pereira, A.G.; Rubira, A.F.; Muniz, E.C. Polyelectrolyte complexes based on pectin–NH2 and chondroitin sulfate. Carbohydr. Polym. 2012, 87, 1950–1955. [Google Scholar] [CrossRef] [Green Version]
  49. Zhang, S.; Zhang, Y.; Fu, L.; Jing, M. A chitosan fiber as green material for removing Cr(VI) ions and Cu(II) ions pollutants. Sci. Rep. 2021, 11, 22942. [Google Scholar] [CrossRef] [PubMed]
  50. Cui, Z.; Xiang, Y.; Si, J.; Yang, M.; Zhang, Q.; Zhang, T. Ionic interactions between sulfuric acid and chitosan membranes. Carbohydr. Polym. 2008, 73, 111–116. [Google Scholar] [CrossRef]
  51. Bouchet, R.; Siebert, E. Proton conduction in acid doped polybenzimidazole. Solid State Ion. 1999, 118, 287–299. [Google Scholar] [CrossRef]
  52. Miraboutalebi, S.M.; Nikouzad, S.K.; Peydayesh, M.; Allahgholi, N.; Vafajoo, L.; McKay, G. Methylene blue adsorption via maize silk powder: Kinetic, equilibrium, thermodynamic studies and residual error analysis. Process. Saf. Environ. Prot. 2017, 106, 191–202. [Google Scholar] [CrossRef]
  53. Jembere, A.L.; Genet, M.B. Comparative adsorptive performance of adsorbents developed from sugar industrial wastes for the removal of melanoidin pigment from molasses distillery spent wash. Water Resour. Ind. 2021, 26, 100165. [Google Scholar] [CrossRef]
  54. Lin, B.; Hua, M.; Zhang, Y.; Zhang, W.; Lv, L.; Pan, B. Effects of organic acids of different molecular size on phosphate removal by HZO-201 nanocomposite. Chemosphere 2017, 166, 422–430. [Google Scholar] [CrossRef]
  55. Li, R.; Wang, J.J.; Zhou, B.; Awasthi, M.K.; Ali, A.; Zhang, Z.; Lahori, A.H.; Mahar, A. Recovery of phosphate from aqueous solution by magnesium oxide decorated magnetic biochar and its potential as phosphate-based fertilizer substitute. Bioresour. Technol. 2016, 215, 209–214. [Google Scholar] [CrossRef] [Green Version]
  56. Shen, H.; Wang, Z.; Zhou, A.; Chen, J.; Hu, M.; Dong, X.; Xia, Q. Adsorption of phosphate onto amine functionalized nano-sized magnetic polymer adsorbents: Mechanism and magnetic effects. RSC Adv. 2015, 5, 22080–22090. [Google Scholar] [CrossRef]
  57. Jung, K.W.; Lee, S.; Lee, Y.J. Synthesis of novel magnesium ferrite (MgFe2O4)/biochar magnetic composites and its adsorption behavior for phosphate in aqueous solutions. Bioresour. Technol. 2017, 245, 751–759. [Google Scholar] [CrossRef] [PubMed]
  58. Lai, L.; Xie, Q.; Chi, L.; Gu, W.; Wu, D. Adsorption of phosphate from water by easily separable Fe3O4@SiO2 core/shell magnetic nanoparticles functionalized with hydrous lanthanum oxide. J. Colloid Interface Sci. 2016, 465, 76–82. [Google Scholar] [CrossRef] [PubMed]
  59. Trinh, V.T.; Nguyen, T.M.P.; Van, H.T.; Hoang, L.P.; Nguyen, T.V.; Ha, L.T.; Vu, X.H.; Pham, T.T.; Quang, N.V.; Nguyen, X.C. Phosphate Adsorption by Silver Nanoparticles-Loaded Activated Carbon derived from Tea Residue. Sci. Rep. 2020, 10, 3634. [Google Scholar] [CrossRef] [Green Version]
  60. Su, Y.; Yang, W.; Sun, W.; Li, Q.; Shang, J.K. Synthesis of mesoporous cerium–zirconium binary oxide nanoadsorbents by a solvothermal process and their effective adsorption of phosphate from water. Chem. Eng. J. 2015, 268, 270–279. [Google Scholar] [CrossRef]
  61. Huang, W.; Chen, J.; He, F.; Tang, J.; Li, D.; Zhu, Y.; Zhang, Y. Effective phosphate adsorption by Zr/Al-pillared montmorillonite: Insight into equilibrium, kinetics and thermodynamics. Appl. Clay Sci. 2014, 104, 252–260. [Google Scholar] [CrossRef]
  62. Deng, Z.; Gu, S.; Cheng, H.; Xing, D.; Twagirayezu, G.; Wang, X.; Ning, W.; Mao, M. Removal of Phosphate from Aqueous Solution by Zeolite-Biochar Composite: Adsorption Performance and Regulation Mechanism. Appl. Sci. 2022, 12, 5334. [Google Scholar] [CrossRef]
  63. Shukla, S.K.; Pandey, S.; Saha, S.; Singh, H.R.; Mishra, P.K.; Kumar, S.; Jha, S.K. Removal of crystal violet by Cu-chitosan nano-biocomposite particles using Box–Behnken design. J. Environ. Chem. Eng. 2021, 9, 105847. [Google Scholar] [CrossRef]
  64. Hidayat, E.; Khaekhum, S.; Yonemura, S.; Mitoma, Y.; Harada, H. Biosorption of Eriochrome Black T Using Exserohilum rostratum NMS1.5 Mycelia Biomass. J 2022, 5, 427–434. [Google Scholar] [CrossRef]
Figure 1. (a) pHzpc of CS-ZL/ZrO/Fe3O4, (b) XRD spectra of CS-ZL/ZrO/Fe3O4, and (c) photograph of CS-ZL/ZrO/Fe3O4 (taken by phone).
Figure 1. (a) pHzpc of CS-ZL/ZrO/Fe3O4, (b) XRD spectra of CS-ZL/ZrO/Fe3O4, and (c) photograph of CS-ZL/ZrO/Fe3O4 (taken by phone).
Ijms 24 09754 g001
Figure 2. SEM images before (a), and after (b) PO43− adsorption.
Figure 2. SEM images before (a), and after (b) PO43− adsorption.
Ijms 24 09754 g002
Figure 3. FTIR-ATR before, and after PO43− adsorption.
Figure 3. FTIR-ATR before, and after PO43− adsorption.
Ijms 24 09754 g003
Figure 4. WR of CS-ZL/ZrO/Fe3O4. (a) Percentage WR, and (b) FTIR-ATR of CS-ZL/ZrO/Fe3O4 after treatment. Solution a (0.01 M HCl), b (0.1 M HCl), c (0.01 M H2SO4), and d (0.1 M H2SO4). Standard deviation (error bars).
Figure 4. WR of CS-ZL/ZrO/Fe3O4. (a) Percentage WR, and (b) FTIR-ATR of CS-ZL/ZrO/Fe3O4 after treatment. Solution a (0.01 M HCl), b (0.1 M HCl), c (0.01 M H2SO4), and d (0.1 M H2SO4). Standard deviation (error bars).
Ijms 24 09754 g004
Figure 5. (a) Pareto chart for the standardized effect of various factors on PO43− removal by adsorbent, (b) pH and dosage of adsorbent response surface’s effect on PO43− removal (%), (c) pH and time response surface’s effect on PO43− removal (%), and (d) pH and dosage of adsorbent and time response surface’s effect on PO43− removal (%).
Figure 5. (a) Pareto chart for the standardized effect of various factors on PO43− removal by adsorbent, (b) pH and dosage of adsorbent response surface’s effect on PO43− removal (%), (c) pH and time response surface’s effect on PO43− removal (%), and (d) pH and dosage of adsorbent and time response surface’s effect on PO43− removal (%).
Ijms 24 09754 g005
Figure 6. Effect of initial concentration on PO43− removal onto CS-ZL/ZrO/Fe3O4. Standard deviation (error bars).
Figure 6. Effect of initial concentration on PO43− removal onto CS-ZL/ZrO/Fe3O4. Standard deviation (error bars).
Ijms 24 09754 g006
Figure 7. Linear curves of PO43− adsorption isotherm models. (a) Langmuir, and (b) Freundlich models. Standard deviation (error bars).
Figure 7. Linear curves of PO43− adsorption isotherm models. (a) Langmuir, and (b) Freundlich models. Standard deviation (error bars).
Ijms 24 09754 g007
Figure 8. The effect of contact time on PO43− removal onto CS-ZL/ZrO/Fe3O4. Standard deviation (error bars).
Figure 8. The effect of contact time on PO43− removal onto CS-ZL/ZrO/Fe3O4. Standard deviation (error bars).
Ijms 24 09754 g008
Figure 9. Linear curves of PO43− adsorption kinetic studies. (a) Pseudo-first-order (PFO) and (b) pseudo-second-order (PSO) models. Standard deviation (error bars).
Figure 9. Linear curves of PO43− adsorption kinetic studies. (a) Pseudo-first-order (PFO) and (b) pseudo-second-order (PSO) models. Standard deviation (error bars).
Ijms 24 09754 g009
Figure 10. The effect of coexisting ions on PO43− removal onto CS-ZL/ZrO/Fe3O4. Standard deviation (error bars). A: no significant effect (p ≤ 0.05).
Figure 10. The effect of coexisting ions on PO43− removal onto CS-ZL/ZrO/Fe3O4. Standard deviation (error bars). A: no significant effect (p ≤ 0.05).
Ijms 24 09754 g010
Figure 11. The percentage of desorption. (a) Different NaOH concentrations and (b) different contact times using 1 M NaOH, and (c) in recycle studies on PO43− adsorption capacity. Standard deviation (error bars).
Figure 11. The percentage of desorption. (a) Different NaOH concentrations and (b) different contact times using 1 M NaOH, and (c) in recycle studies on PO43− adsorption capacity. Standard deviation (error bars).
Ijms 24 09754 g011
Table 1. Experimental data results from 4 factors of BBD for PO43− removal onto CS-ZL/ZrO/Fe3O4.
Table 1. Experimental data results from 4 factors of BBD for PO43− removal onto CS-ZL/ZrO/Fe3O4.
RunpHDosageTemperatureTime% Removal
120.02453558.95
2100.02453551.75
320.10453572.77
4100.10453554.54
570.06301061.83
670.06601056.68
770.06306064.95
870.06606064.94
920.06451059.09
10100.06451054.34
1120.06456072.75
12100.06456059.57
1370.02303556.02
1470.10303561.46
1570.02603556.69
1670.10603559.69
1720.06303566.84
18100.06303560.16
1920.06603564.70
20100.06603555.08
2170.02451054.40
2270.10451058.87
2370.02456056.72
2470.10456069.58
2570.06453553.48
2670.06453555.47
2770.06453553.47
Table 2. Physical properties of the adsorbent.
Table 2. Physical properties of the adsorbent.
Specific Surface AreaValue
BET-specific surface area (m2/g)88.1
Pore volume (mL/g)0.572
Average diameter (µm)43.9
Porosity (%)59
Table 3. EDS data before and after PO43− adsorption.
Table 3. EDS data before and after PO43− adsorption.
ElementWeight %Atomic %
BeforeAfterBeforeAfter
N3.272.0613.437.66
Al0.780.931.651.79
Si6.351.7012.983.15
Fe38.9256.3940.0352.69
Zr50.6827.7631.9115.88
P 11.17 18.82
Table 4. ANOVA results for PO43− removal.
Table 4. ANOVA results for PO43− removal.
SourceDFSum of SquaresMean of SquaresF-Valuep-ValueRemarks
Model14839.75159.98216.68<0.0001 *Significant
A1296.610296.61051.8093<0.0001 *
B1149.672149.67212.5582<0.0001 *
C115.14315.1431.28350.063
D1156.241156.2410.6342<0.0001 *
A2174.28574.28512.81510.001 *
B219.6429.6421.47220.127
C2167.00867.0087.12790.001 *
D2190.95290.9521.9374<0.0001 *
A × B130.41530.4154.27340.013 *
A × C12.1612.1610.30360.453
A × D117.76617.7668.21180.046 *
B × C11.4881.4880.20910.532
B × D117.59817.5980.05400.047 *
C × D16.6056.6051.840.200
Error1243.1573.596
Lack-of-Fit1040.5044.0503.050.272
Pure Error22.6531.3267
Total26882.908
R2 95.11
R2 adj 89.41
* Significant.
Table 5. Isotherm model parameters for PO43− removal onto CS-ZL/ZrO/Fe3O4.
Table 5. Isotherm model parameters for PO43− removal onto CS-ZL/ZrO/Fe3O4.
IsothermsParametersValue
Langmuirqmax1259.79
KL14.27
R20.7409
RL0.0007
FreundlichKf1135.07
1/n0.7555
R20.9970
Table 6. Kinetic model parameters for PO43− removal onto CS-ZL/ZrO/Fe3O4.
Table 6. Kinetic model parameters for PO43− removal onto CS-ZL/ZrO/Fe3O4.
KineticsParametersValue
PFOqe2.5165
K11.42857 × 10−6
R21.00 × 10−4
PSOqe510,204.1
K20.000119
R20.9979
Table 7. List comparing PO43− adsorption amounts.
Table 7. List comparing PO43− adsorption amounts.
AdsorbentpHqe (mg/g)References
Magnetic iron oxide nanoparticles115.03[1]
Fe-HNT45.46[18]
Halloysite43.56[18]
20MMSB4121.25[55]
Amine-functionalized nano magnetic Fe3O4 polymer3.0102.04[56]
MFB-MCs3.0487.99[57]
Fe3O4@SiO2 core/shell magnetic nanoparticles-27.8[58]
AgNPs-TAC313.62[59]
Ce0.8Zr0.2O26.2112.23[60]
Zr/Al-Mt5.017.2[61]
PZC 7.3112.41[62]
Zeolite110.69[62]
Biochar113.60[62]
CS-ZL/ZrO/Fe3O42732.56Present study
Table 8. Variables and levels.
Table 8. Variables and levels.
SymbolFactorLevel 1 (− 1)Level 2 (0)Level 3 (+ 1)
ApH2710
BDosage (g)0.020.060.10
CTemperature (°C)304560
DTime (min)103560
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hidayat, E.; Mohamad Sarbani, N.M.B.; Yonemura, S.; Mitoma, Y.; Harada, H. Application of Box–Behnken Design to Optimize Phosphate Adsorption Conditions from Water onto Novel Adsorbent CS-ZL/ZrO/Fe3O4: Characterization, Equilibrium, Isotherm, Kinetic, and Desorption Studies. Int. J. Mol. Sci. 2023, 24, 9754. https://doi.org/10.3390/ijms24119754

AMA Style

Hidayat E, Mohamad Sarbani NMB, Yonemura S, Mitoma Y, Harada H. Application of Box–Behnken Design to Optimize Phosphate Adsorption Conditions from Water onto Novel Adsorbent CS-ZL/ZrO/Fe3O4: Characterization, Equilibrium, Isotherm, Kinetic, and Desorption Studies. International Journal of Molecular Sciences. 2023; 24(11):9754. https://doi.org/10.3390/ijms24119754

Chicago/Turabian Style

Hidayat, Endar, Nur Maisarah Binti Mohamad Sarbani, Seiichiro Yonemura, Yoshiharu Mitoma, and Hiroyuki Harada. 2023. "Application of Box–Behnken Design to Optimize Phosphate Adsorption Conditions from Water onto Novel Adsorbent CS-ZL/ZrO/Fe3O4: Characterization, Equilibrium, Isotherm, Kinetic, and Desorption Studies" International Journal of Molecular Sciences 24, no. 11: 9754. https://doi.org/10.3390/ijms24119754

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