Next Article in Journal
Preparation of a Hydrogel Nanofiber Wound Dressing
Next Article in Special Issue
Facile Synthesis of Carbon- and Nitrogen-Doped Iron Borate as a Highly Efficient Single-Component Heterogeneous Photo-Fenton Catalyst under Simulated Solar Irradiation
Previous Article in Journal
Dual Plasmon Resonances and Tunable Electric Field in Structure-Adjustable Au Nanoflowers for Improved SERS and Photocatalysis
Previous Article in Special Issue
Advanced Oxidation Processes Coupled with Nanomaterials for Water Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Fe3O4@mZrO2-Re (Re = Y/La/Ce) by Using Uniform Design, Surface Response Methodology, and Orthogonal Design & Its Application for As3+ and As5+ Removal

1
Engineering Research Center of Ministry of Education for Mine Ecological Restoration, China University of Mining and Technology, Xuzhou 221116, China
2
Department of Biological Sciences, Karakoram International University, Gilgit 15100, Pakistan
3
Department of Environmental Science, Karakoram International University, Gilgit 15100, Pakistan
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(9), 2177; https://doi.org/10.3390/nano11092177
Submission received: 13 July 2021 / Revised: 21 August 2021 / Accepted: 21 August 2021 / Published: 25 August 2021

Abstract

:
In this study, iron oxide (Fe3O4) was coated with ZrO2, and doped with three rare earth elements((Y/La/Ce), and a multi-staged rare earth doped zirconia adsorbent was prepared by using uniform design U14, Response Surface methodology, and orthogonal design, to remove As3+ and As5+ from the aqueous solution. Based on the results of TEM, EDS, XRD, FTIR, and N2-adsorption desorption test, the best molar ratio of Fe3O4:TMAOH:Zirconium butoxide:Y:La:Ce was selected as 1:12:11:1:0.02:0.08. The specific surface area and porosity was 263 m2/g, and 0.156 cm3/g, respectively. The isothermal curves and fitting equation parameters show that Langmuir model, and Redlich Peterson model fitted well. As per calculations of the Langmuir model, the highest adsorption capacities for As3+ and As5+ ions were recorded as 68.33 mg/g, 84.23 mg/g, respectively. The fitting curves and equations of the kinetic models favors the quasi second order kinetic model. Material regeneration was very effective, and even in the last cycle the regeneration capacities of both As3+ and As5+ were 75.15%, and 77.59%, respectively. Adsorption and regeneration results suggest that adsorbent has easy synthesis method, and reusable, so it can be used as a potential adsorbent for the removal of arsenic from aqueous solution.

Graphical Abstract

1. Introduction

Arsenic is the 20th most toxic and carcinogenic metallic element found in nature, and its content in the earth crust is about 2–5 mg/kg [1]. People are exposed to arsenic mainly by the direct contact with drinking water and indirect contact through food chain transmissions. When arsenic and its compounds are ingested by the human body, they can accumulate in human body, and the main parts of human body that are prone to arsenic accumulation are hair, nails, bones, liver, kidney, etc. [2]. If the human body is directly exposed to a large amount of arsenic, the central nervous system transmission will be diminished, resulting in numbness of hands and feet, gastrointestinal and respiratory tract lesions etc. [3]. It is proven that arsenic is usually found in the form of arsenic oxides and arsenic trioxides. Inorganic arsenic is divided into As(III) and As(V). The toxicity level of arsenic(III) is higher than arsenic(V) [4]. Therefore, it is crucial to control arsenic pollution in the environment, especially in the water environment.
The most common methods used for arsenic removal include; precipitation, coagulation and sedimentation, electro-dialysis, electro-coagulation, reverse osmosis, nanofiltration, ion-exchange, oxidation, adsorption, etc. [3]. These technologies have their advantages, disadvantages, and conditions of use. There are also differences in the operating costs, simple usage, and long-term operational reliability [5,6]. Among these methods, adsorption is one of the most promising approaches for the treatment of arsenic polluted wastewater [7]. In the last many years, different potential adsorbents like hydroxides, activated carbon, graphene, activated alumina etc. have been synthesized, and used for arsenic removal [8]. Zirconia is one of the adsorbents used for arsenic removal and many studies suggests that Zirconia has a good adsorption effect for arsenic removal from the aqueous solutions. The ZrO2 layer can rise the adsorption properties of adsorbents to achieve the efficient removal of As(III) and As(V) from the water [9]. Rare earth doped nanomaterials have been synthesized and extensively used for water pollution, disease diagnosis, drug delivery, biocompatibility, drug loading etc. [10].
Uniform design is a statistical tool used for examining the relationship between various experimental variables to one or more responses [11]. Uniform design is particularly well suited to multi-factors, and multi-level assessments, such as assessing nanomaterial preparation molar ratios [12,13]. Response surface methodology (RSM) can give a spontaneous graph to impulsively observe the optimization points, and spontaneously determine the optimization areas [14]. Uniform Design and RSM together are very effective in optimizing the synthesis ratios of magnetic nanomaterials [15,16]. The orthogonal experiment method is a type of design method used to investigate a variety of factors and levels. It performs tests by choosing a suitable number of representative test cases from a large set of test data [17]. Although there are many studies on adsorption of arsenic have been reported. However, there are very limited or no studies on material optimizations using RSM, Uniform design, and also orthogonal design and multi doping of rare earth metals.
In this paper by using uniform design U14*(145) and response surface analysis (RSM), the synthesis ratio of zirconium oxide was determined; by using L9 orthogonal design table the doping ratio of rare earth elements was determined. Where, iron ixide (Fe3O4) was coated with ZrO2 and doped with three rare earth elements (Y/La/Ce), so to prepare rare earth doped zirconium adsorbent to treat As3+ and As5+ from water. Through different characterization analysis like XRD, FT-IR, N2 adsorption-desorption, EDS, VSM, and TEM, the morphology, elemental composition, pore size, specific surface area, and crystallinity were determined. Then adsorption efficiencies of As3+ and As5+ on to the prepared material was examined.

2. Materials and Methods

2.1. Experimental Materials

China Pharmaceutical Group Chemical Reagents Co., Ltd. (Shanghai, China) supplied FeCl3-6H2O of Analytically pure grade. Analytically Pure Sodium Acetate was received from Xilong Science Co., Ltd. (Shantou, China). China Pharmaceutical Group Chemical Reagents Co., Ltd. provided pure Analytical Grade Nitric Acid (HNO3), Sodium Arsenate (Na3AsO4-12H2O), and Polyvinylpyrrolidone K30 (PVP. Absolute Ethanol (CH3CH2OH) was received from Shanghai Zhongqin Chemical Reagent Co., Ltd. (Shanghai, China). Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China) provided Industrial grade Zirconium (IV) butoxide (Zr(OBu)4). Guangdong Weng Jiang Chemical Reagent Co., Ltd. (Shaoguan, China) produced analytically pure lanthanum nitrate hexahydrate La(NO3)3·6H2O, and Cerium nitrate hexahydrate Ce(NO3)3·6H2O. Analytically Pure Sodium Arsenite (NaAsO2) was purchased from Chemical Reagent Research Center of Guangdong (Guangzhou, China).

2.2. Characterization Methods

The micro-morphology of the sample was observed and examined using an American FEI Tecnai G2 F20 Field Transmission Electron Microscope (TEM), at 200 kV acceleration voltage. VSM was conducted by using a Vibrating Sample Magnetometer (Quantum Design, PPMS-9, San Diego, CA, USA) at a room temperature under the temperature-controlled conditions in the range of 2–400 K. A qualitative and quantitative study of the surface elemental composition of samples was carried out using an FEI Quanta FEG 250 Scanning Electron Microscope from the United States in the range of 200 V–30 kV electron beam voltage. FT-IR Spectrometer (VERTEX 80V, Brunker, Germany) was used for the characterization of groups on material surface. The sample was dried and dispersed in KBr powder and passed into pellets. Spectrum range: 4000–400 cm−1; resolution: 0.06 cm−1; beam diameter 40 mm. The The X-ray powder diffraction (XRD) patterns were collected by using X’PERT powder small angle X-ray diffractometer (PANalytical, Almelo, The Netherlands). The X-ray radiation source was a ceramic X-ray Diffraction Copper anode (Cu Kα radiation: λ = 1.54056 Å) and Bragg-Brentano configuration, and diffractions were taken at room temperature.

2.3. Synthesis of Fe3O4 and Fe3O4@mX-ZrO2, X = Y/La/Ce

Iron oxide (Fe3O4) was prepared using our previously reported work. 80 mL Ethylene Glycol, 4.50 g FeCl3-6H2O and 10g CH3COONa were first added in a round bottom flask, and mixed it well. 3.2 g TMAOH was separately dissolved in 20 mL Ethylene glycol and then added dropwise to the main solution, which was then mixed for 30 min using magnetic shaking. The solution was then transferred to a Teflon-lined stainless steel Autoclave and heated at 200 °C for 8 h. After cooling down at room temperature, the prepared MNPs were then cleaned several times with alcohol and deionized water, and dried in an oven at 50 °C for 12 h [16]. Then Fe3O4@mX-ZrO2, X = Y/La/Ce was prepared, where 1 g of PVP was first weighed in 20 mL of deionized water and after ultrasonic dispersion, a certain amount of magnetic Fe3O4 was added, and again the ultrasonic dispersion continued for 30 min. Then, the required amount of TMAOH was added, and continued ultrasonic shaking for 10 min. Meanwhile, 10mL ethanol, zirconium butoxide, rare earth elements (Y/La/Ce) were added into small glass vials and ultrasonically shook for 10 min. Then, these were added dropwise to the main solution of the system, and again ultrasonic shaking for 30 min, later solution was transferred to a three-mouth round bottom flask, and stimulated at room temperature for 3 h at 200 r/min. Subsequently, the solution was transferred to an oven and heated at 150 °C for 3 h for the hydrothermal reaction. Lastly, after cooling down, material was washed with alcohol and deionized water, and dried in the vacuum oven at 60 °C for 12 h.

2.4. Optimizing the Synthesis Ratio of Fe3O4@mZrO2-Re (Re = Y/La/Ce)

2.4.1. Uniform Design

Uniform design is the statistical technique mostly used to study the relationship between different experimental factors to one or many related responses. In most of these cases, to run a complete factorial design to get sufficient resources may not always exist, so, therefore, small factorial designs are frequently used to significantly reduce the experiment numbers [11,12]. The uniform design table, the supporting tables, and the seven-level design are shown in Table 1, Table 2, Table 3 and Table 4. In Table 1 independent variables and experiment sequence is arranged and selected.
The rules for creating the columns of U14*(145) are given in Table 2 and on the basis of this further designing of the selected factors was achieved.
By using the Table 1 and Table 2, seven levels of the three selected variables i.e., Fe3O4 dosage (X1), TMAOH dosage(X2), and zirconium butoxide dosage (X3) were derived, and laid a foundation for further use of uniform design. The seven levels of the selected variables are listed in Table 3.
The regression analysis model formula was used for analyzing the influencing factors of each variable, and to estimate, evaluate, and optimize the experiment conditions. The autoregressive-response model was fitted with quadratic polynomials since this model clearly expresses the interaction between independent variables. The fitting method is stepwise regression method (Equation (1)). Specific experimental arrangements and results are shown in Table 4.
                      Y = β 0 + i = 1 m β i X i + i = 1 m β i j X i 2 + i < j m β i j X i X j + ε
Based on the designed U14 uniform experiment design, the results for each level was measured and identified the best combinations for adsorption. The amount of As5+ adsorbed on material surface at equilibrium was the response value; each factor was set at seven levels to investigate the relationship between factors and their impact on response values.

2.4.2. RSM Analysis

RSM can determine the effects of various factors and their relations to the investigated indicators (response values) in scientific experiments. It can also be used to fit a comprehensive quadratic polynomial model through a fundamental extravagant experiment, which can give better experimental design and results [18]. For RSM, the main selected factors were; factor-1: Fe3O4, factor-2: TMAOH, and factor-3: Zirconium(IV) Butoxide, and the main factor chosen for this was Zirconium(IV) Butoxide.

2.4.3. Orthogonal Design

To study the method of multi-factor and multi-level design, and to find the optimal level combination through a small number of experiments, orthogonal experiment design can be used. The larger the number and level of factors, the clearer the benefits of this method. Orthogonal experiment design is an efficient, fast, and economic experiment design method based on the orthogonal table [19]. To calculate the doping amount of the rare earth elements orthogonal design was applied. The arrangement of every factor and levels were optimized by using the orthogonal design of L9. The optimum ratio of the selected doping elements for the preparation of doped material were, A (Yttrium), B (Lanthanum), C (Cerium), and D (empty list), and the dosage of every cube of doping material in parentheses.

2.5. Adsorption Procedure

By using Sodium Arsenite (NaAsO2) and Sodium arsenate (Na3AsO4·12H2O) stock solutions of As3+, and As5+ were prepared and used for the adsorption experiments. The standard solutions were gradually diluted by using the stock solutions. The accurate weighing 0.01 g adsorbent was put into a 250 mL conical flask with a stopper and added 100 mL solutions of As3+, and As5+. The solution was adapted to the appropriate pH value by using HCL and NaOH. The tapered flask was then put in a persistent temperature oscillator at 200 r/min for 60 min. After one hour of oscillation, magnetic separation was used, and the supernatant was filtered by a membrane (0.45 μm), and concentration of the heavy metals was detected by ICP. The formula for adsorption capacity is as follows.
Q e = C 0 C e × ν m
where; Qe (mg/g) is the adsorption equilibrium capacity, C0 (mg/L) is the initial ion concentration, Ce (mg/L) after adsorption ion concentration in solution, V (L) is the volume of the solution to be adsorbed, L; and m (g) is the adsorbent dosage.

3. Results and Discussions

3.1. Use Uniform Design, RSM and Orthogonal Design for Optimizing Adsorbent

3.1.1. Uniform Design

On the basis of guidelines in the Table 1, Table 2 and Table 3 of the uniform design, the selected variables and their concentrations at designed 7 levels were calculated, and data is shown in the Table 4. X1, X2, and X3 were selected as the independent variables, which were Fe3O4 dosage, TMAOH dosage, and zirconium butoxide dosage, respectively. For each adsorption level, As(V) concentration was taken as 5 mg/L, volume 100 mL, adsorbent 0.01 gm, adsorbed at room temperature for 60 min, and performed the experiments at room temperature.
To get the adsorption capacity model of the prepared adsorbent for As5+ removal, the DPS-7.05 (Data Processing System) was used to choose the quadratic polynomial stepwise regression method to achieve regression analysis. The data processing system (DPS) is a comprehensive and user-friendly platform for experimental designs, statistical analysis, and data mining that provides uniform design and calculation benefits [20]. Equation (3) was used to examine the final values of Y‘, which are the predicted values of equilibrium adsorption capacities, and results are mentioned in Table 4.
Y = 2.762519818 + 179.02323398 X 3 + 235.52925909 X 4 5632.701746 X 1 2 35731.654194 X 3 2 4378.245969 X 4 2 8720.569751 X 1 × X 2 + 2660.3369935 X 1 × X 4 + 12910.796840 X 2 × X 3
Fisher variance is a substantial tool to find the differences of means between two or more samples. In order to investigate the differences between the experimental and predicted values, Fisher’s variance test was used. If compared with given standard F-value, the larger the F-value is, the clearer the treatment result, and also test precision is higher [16]. Uniform design samples were prepared according to the range and level listed in the Table 4. The variance test of Experimental values (Y) and predicted values (Y‘) of the samples are shown in Figure 1.
With variance test, the Fisher variance test F-value is 13.38, which is much higher than F(0.05,8,5) = 4.81. The correlation coefficient R2 is 0.9683, which is in good agreement with the experimental value Y, showing that only 3.17% total variance could not be explained by the model. Therefore, Equation (3) can be used to predict the molar ratio of the samples.

3.1.2. RSM Analysis

The factors selected for RSM analysis were; Factor-1: Fe3O4, Factor-2: TMAOH, and Factor-3: Zirconium(IV) butoxide (Zr(OBu)4), and Zirconium(IV) butoxide were selected as the main factor. The Figure 2 and Figure 3 explains the model and methodology used to find the best adsorption combinations for As5+ removal from the aqueous solution.
From Figure 2 can be noticed that the response surface exhibits a clear frontal. The adsorption response surface of As5+ is curved with the change of Zr(OBu)4 dosage, which indicates that it is a vital factor to affect the adsorption process. With the increase in Zr(OBu)4 dosage, the adsorption also increased, but after sometimes it decreases, so an appropriate amount of Zr(OBu)4 dosage improved the adsorption [15]. From the contours of the interaction between Zr(OBu)4 and Fe3O4, it can be noticed that the high-value region appeared in the middle region i.e., Zr(OBu)4: 0.005~0.0058 mmol, and Fe3O4: 0.00048~0.00055 mmol.
From Figure 3 it can be noted that the adsorption response surface of As5+ is curved with the change of Zr(OBu)4 and TMAOH dosage, which indicates that both are crucial factors to affect the adsorption effect. From the contours of the interaction between TMAOH and Zr(OBu)4, it can be noticed that the high-value region appeared in the middle region (TMAOH: 0.0058~0.0065 mmol and Zr(OBu)4: 0.005~0.0058 mmol).

3.1.3. Orthogonal Design

Based on the orthogonal design As(III) adsorption results the best doping combinations for adsorbent were finalized. The molar ratio of Fe3O4:TMAOH:Zr(OBu)4 for this design was kept at 1:12:11. The orthogonal design arrangement of each rare earth element (La/Ce/Y) and level is given in Table 5.
The revealed data from the Orthogonal design L9 (Table 5) suggests that the excellent combination was noted at level A2B1C2. Based on analyzed adsorption results, the optimal mixture ratio for the doping combination of the rare earth metals (Y/La/Ce) to prepare the best doping combinations for the removal of As(V) was finalized, which was as follows; Y: 1%; La: 0.2%; Ce: 0.8%. The relative molar ratio of Fe3O4:TMAOH:Zirconium(IV) butoxide was determined as 1:12:11, while the final best doping combination was found to be Y: 1%: La: 0.2%: Ce: 0.8. The best molar ratio for Fe3O4:TMAOH:Zr(OBu)4:Y:La:Ce was achieved as 1:12:11:1:0.02:0.08.

3.2. Characterizations of Fe3O4@mZrO2-Re (Re = Y/La/Ce)

3.2.1. TEM Analysis

The morphology of material before and after surface modifications was tested by Transmission election microscopy (TEM). Figure 4 shows the TEM images of Fe3O4 (a), and Fe3O4@mZrO2-Re (Re = Y/La/Ce) (b). Uncoated Fe3O4 were observed as dispersed spherical particles with the diameter ranging 0.032–0.065 um. The good dispersibility probably originates from the TMAOH that effectively prevented from the agglomeration [21]. By contrast, the surface of Fe3O4@mZrO2-Re (Re = Y/La/Ce) exhibited uneven and larger morphologies probably due to the coating with ZrO2 and doped with rare earth metals.

3.2.2. VSM Analysis

Magnetic saturation intensity is a key factor to check the magnetization of any magnetic material. The magnetic characteristics of the materials was investigated by Vibrant Sample Magnetometer (VSM) at room temperature. The magnetic hysteresis loops of the Fe3O4 and Fe3O4@mZrO2-Re (Re = Y/La/Ce) mesoporous materials are shown in Figure 5. As shown in the Figure 5 that the saturation magnetization values of Fe3O4 and Fe3O4@mZrO2-Re (Re = Y/La/Ce) are 49 emu/g and 19.8 emu/g, respectively. It can be observed that after the coating with ZrO2 and doping with rare earth metals, the magnetic saturation intensity significantly reduced i.e., from 49 emu/g to 19.8 emu/g. The decrease in the magnetic intensity saturation may be because of the addition of layers on the surface of the Fe3O4 but still, the magnetization value was enough for the magnetic separation in a short time by using an external magnet. The same kind of results are also described in other related studies [16,22].

3.2.3. EDS Analysis

The Energy Dispersive Spectroscopy (EDS) of Fe3O4@mZrO2-Re (Re = Y/La/Ce) was conducted to determine the constituent elements, and corresponding peaks are shown in Figure 6, and elemental composition is listed in Table 6.
The EDS results in Figure 6 and Table 6 confirmed the presence of Fe, O, Zr, Ce, Y, and La, indicating that the prepared material have successfully coated with zirconia and doped with rare earth elements on the surface of Fe3O4. Table 6 shows that the weight% of Yttrium (Y), Cerium (Ce), and Lanthanum (La) is zero before doping, while it shows as 1.08%, 1.02%, and 0.45%, respectively, after doping, which confirms the doping.

3.2.4. XRD Analysis

The XRD patterns of Fe3O4@mZrO2-Re (Re = Y/La/Ce) are shown in Figure 7. The XRD profile with weak signals towards the halo peak probably indicate that the magnetite contains both crystalline and amorphous parts [23]. The diffraction peaks at 2θ = 30.4, 35.8, 38.2, and 43, are assignable to the 220, 311, 200, and 400 Lattice planes of face centered-cubic magnetite (JCPDS No.19-0629), respectively [22].

3.2.5. FTIR Analysis

FT-IR (Fourier transform infrared spectra) characterization was done to detect the major structural groups present in Fe3O4@mZrO2-Re (Re = Y/La/Ce). FTIR spectra of the material is shown in Figure 8.
As shown in Figure 8 that the FTIR spectrum of Fe3O4@mZrO2-Re (Re = Y/La/Ce) NPs stretches from 4000–400 cm−1, where the FTIR peaks at 498 cm−1 are likely attributed to Fe-O Stretching [24]. The characteristic band at 3419 cm−1 and 1623 cm−1 represent the stretching and bending of modes of OH group of the adsorbed water present on the surface of the nanomaterials [25]. The band appeared at 602 cm−1 of the spectra are assigned to the zirconia group. The peaks with obvious fluctuation are all related to hydroxyl groups. The valence band electrons of ZrO2 can be excited under certain conditions, and an electron-hole pair can be formed on its surface. The hole then can react with oxygen ion on the surface of ZrO2 to form an oxygen hole. Under the influence of the polar Zr-O bond, it dissociates to form a surface hydroxyl [26]. However, ZrO2 doped with Y, Ce, and La is likely to enhance the volume of hydroxyl groups at the surface of ZrO2 to upgrade its adsorption capacity.

3.2.6. N2 Adsorption-Desorption Analysis

The porous characteristics of Fe3O4@mZrO2-Re (Re = Y/La/Ce) was examined by nitrogen adsorption isotherm measurement. The specific surface area and pore size distribution of the materials was measured from the analysis of desorption branch of isotherms using the density function theory.
As shown in Figure 9, an isotherm is typical for mesoporous material with a hysteresis loop at partial pressures [27]. According to Brunauer-Emmett-Teller (BET) analysis, the prepared materials exhibited large specific surface area of 265 m2/g, pore volume 0.156 cm3/g, and pore diameters 2–5 nm, respectively. N2 adsorption-desorption isotherms of the materials displayed conventional type IV curves with a sharp uptake at a high relative pressure, which demonstrates the existence of cavities between particles.

3.3. Adsorption Performance of Fe3O4@mZrO2-Re (Re = Y/La/Ce)

3.3.1. Effect of pH

The initial pH value of the solution has a significant impact on metal ion adsorption. This parameter is directly related to the capability of hydrogen ions and metal ions to adsorb on the active surface sites [28]. Therefore, the effect of initial pH value on the adsorption performance of Fe3O4@mZrO2 (Re = Y/La/Ce) was investigated. 10 mg of adsorbent was accurately measured and put into different conical flasks, and added 100 mL of 5 mg/L solutions of As3+ and As5+, and adjusted the pH to 3.0, 4.0, 5.0, 6.0, 7.0, 8.0 and 9.0 by using HCL and NaOH. The remaining process was done by following the method in the Section 2.5.
From the Figure 10, it can be clearly noticed that the initial pH value of the solution has a substantial effect on the adsorption. Under the strong basic conditions, the adsorption capacity of the adsorbent for As3+ and As5+ removal was extremely low, while under weak acidic conditions, the adsorption capacities of both metals were high. The highest adsorption capacities for both As3+ and As5+ were recorded at pH 6 and pH 5, respectively. The adsorption capacity of As3+ from pH 5~pH 7 was almost near i.e., 30 mg/g, 32 mg/g and 30 mg/g, respectively. Based on initial pH studies the ideal adsorption pH conditions for both As3+ and As5+ were selected as 6 and 5, respectively.
The relevant studies suggest that that mesoporous zirconia-based nanostructures demonstrated good results for As(V) removal under lower acidic and neutral pH conditions [29]. For as (V), it mainly exists in the form of H2AsO4 and in the range of pH = 3~7. At this time, the adsorbent surface has a positive charge, and the negatively charged H2AsO4 will be adsorbed on the adsorbent surface directly through electrostatic attraction. With the increase of pH, the positive charge on the surface of the adsorbent decreases gradually, so the adsorption capacity of the adsorbent for As5+ decreases gradually; When the pH is greater than 7, the surface of the adsorbent begins to be occupied by negative charges [30]. Some other studies have shown that under acidic conditions, the surface of metal oxides prone to “protonation”, so the adsorption is negatively charged H2AsO4 and HAsO42−, and the protonation decreases with the increase of pH value. The protonation of the sorbent surface is supported by a lower pH. Improved protonation is supposed to escalate the positively charged spots, increase the attraction force between the sorbent surface and As anions, and hence boost adsorption in the lower pH range. The negatively charged sites dominate at higher pH levels, the repulsion effect increases, and the number of adsorption decreases. When the solution is alkaline, OH in the system will compete with arsenate anion for adsorption, resulting in the decrease of arsenic removal efficiency [31]. May be the same inclination was observed with the As3+, although the decline in As3+ removal was not that clear as that of As5+ with the increasing pH.

3.3.2. Adsorption Isothermal Analysis

For non-linear fitting adsorption isothermal analysis, the three widely used isothermal adsorption models are the Langmuir, Freundlich, and Redlich–Peterson models [32]. The specific equation and application scope of each model are as follows:
Langmuir Isothermal non-linear adsorption equation is;
q e = q m K L C e 1 + K L C e
where: qe (mg/g) adsorption capacity at equilibrium, Ce (mg/L) is equilibrium concentration, b (L/mg) is the Langmuir constant, and qm (mg/g) is the Saturated adsorption capacity.
RL is a dimensionless constant separation factor. Its definition formula is;
R L = 1 1 + b C 0
For the adsorption process on the heterogeneous surfaces, Freundlich isotherms are more suitable. The isotherm provides expressions for surface heterogeneity as well as the exponential distribution of active centers and their associated energies. Non-monolayer adsorption is the most common in this method [33]. The equation is as follows:
q e = K F   · C e 1 / n
where: qe (mg/g) is adsorption capacity, KF is the Freundlich coefficient (mg/g (L/mg)1/n), Ce (mg/L) is Equilibrium concentration, and “n” is the Freundlich constant.
The Redlich–Peterson model combines the Langmuir and Freundlich models formula to summarize the Redlich–Peterson isothermal model [34], and the equation is as follows;
q e = K · C e 1 + α · C e β
where: K, α are Redlich–Peterson constant and “β” is coefficient.
The accurately weighted 10 mg adsorbent was put into different 250 mL conical flasks, and 100 mL solutions of As3+, and As5+ with different concentrations like 5 mg/L, 10 mg/L, 20 mg/L, 40 mg/L, 60 mg/L, and 80 mg/L were added, respectively. The solution was adjusted by HCL and NaOH to the required pH value. The conical flasks were then put in the constant temperature oscillator at 298 K, 308 K, and 318 K temperature conditions oscillated at 200 r/min for 60 min. After one hour of oscillation, magnetic separation was used, and the supernatant was then filtered through membrane (0.45 μm), and heavy metals concentration was detected by ICP-OES. According to change in the Ion concentration before and after adsorption the adsorption isotherm was drawn. Figure 11 and Table 7 shows the different adsorption isothermal curves and parameters of the models.
From the Isothermal fitting curves (Figure 11), and equations parameters of the Langmuir model, Freundlich model, and Redlich–Peterson Model, the best adsorption isothermal models for As3+ and As5+ removal was evaluated. The Langmuir model effectively described the adsorption isotherm data with all correlation coefficient (R2) ranged between 0.995~0.997 for As3+, and 0.988~0.990 for As5+. The RL values calculated by the Langmuir model at different temperatures for As3+ and As5+ were ranged from 0.389~0.954, and 0.473~0.531, respectively, all were found less than 1, which suggested that the adsorption is in the favorable direction [35]. The results were found consistent, and the Qm fitted by the Langmuir model is slightly larger than the actual experimental results. The highest adsorption capacity of As3+ ions at 298 K, 308 K, and 318 K, were 62.00 mg/g, 64.31 mg/g, and 68.33 mg/g, respectively, and for As5+ at different temperatures were 59 mg/g, 64 g/g, and 84.23 mg/g, respectively. Based on R2 analysis, the Langmuir model and Redlich–Peterson model shows better results than the Freundlich model, indicating they are best fitted for As3+ and As5+ adsorption isotherm. Based on the analyzed isothermal results of the adsorption isotherms of As3+, and As5+ ions are mainly monolayer layer adsorption and absorption sites on the adsorbent are distributed homogenously [36].
The comparison of the prepared adsorbent Fe3O4@mZrO2-Re (Re = Y/La/Ce) with some relevant studies is shown in Table 8, where the comparison results revealed that the adsorbent is very effective in removing As3+ and As5+ from the aqueous solution.

3.3.3. Adsorption Kinetics of Fe3O4@mZrO2-Re (Re = Y/La/Ce)

The two widely used kinetics models are; (1) the Pseudo-First order kinetic model, a basic kinetic rate equation of stable adsorption reaction; (2) the second one is Pseudo-second order reaction kinetics model, where adsorption rate of adsorbent is directly proportional to the number of available active sites on the surface [45,46,47]. In this study, pseudo-first-order and pseudo-second-order linear models were used to establish the rate constant and As(III) and As(V) controlling mechanism of adsorption onto the prepared adsorbent.
(1)
Pseudo-first-order kinetics model
d q t d t = k 1 q e + q t
Converting the above equation into a linear equation:
Log q e q t = log q e k 1 2.303 t
where: qe (mg/g) is adsorption capacity at equilibrium, qt (mg/g) is the adsorption capacity of materials at time “t”, t (min) is adsorption reaction time, and K1 (min−1) is the first order adsorption rate constant.
(2)
Pseudo-second-order kinetics model
d q t d t   = K 2 q e q t 2
The equation given above is converted into a linear equation:
t q t = 1 K 2 q e 2 + 1 q e t
where: qt (mg/g) is the adsorption capacity at equilibrium, qt (mg/g) adsorption capacity at time “t”, t (min) is the adsorption reaction time, and K1 (min−1) is the second kinetic order adsorption rate constant.
Adsorption kinetic tests were conducted with 0.01 g adsorbent at initial concentrations of (1.0 mg/L, 2.0 mg/L, and 5.0 mg/L), pH 6.0, temperature 298 K, and at time durations of 10, 20, 40, 60, 80, 100, and 120 min. First and second linear kinetic models using Origin 8.5 software were applied to establish the rate constant and As3+ and As5+ adsorption controlling mechanism. The R2 (correlation coefficient) was used to find the most suited model. The values of K1, K2, and qe were also calculated.
From the fitting curves (Figure 12) and related parameters (Table 9) fitted by the kinetic models, it can be noticed that the R2 values obtained from the fittings of the two kinetic reaction models are significantly different. Among the two selected models, Pseudo-second-order kinetic model has a suitable effect, and R2 values were found for both As3+ and As5+ between 0.991–0.997, and found relatively stable, because there is no obvious fluctuation noted. The range of R2 values fitted by the Pseudo-first-order kinetics model are relatively wide ranged for both heavy metals, with values ranging from 0.921 to 0.987, with noticeable fluctuations. Therefore, the adsorption kinetics of As3+ and As5+ from solution by the adsorbent Fe3O4@mZrO2-Re (Re = Y/La/Ce) is more suitably designated by the Pseudo-second-order reaction kinetic model, where rate of reaction depended directly on the square of the concentration of the reactants remaining in the solution [48]. Many other researchers suggested that adsorption kinetics of Magnetic nanomaterials are supporting the Pseudo-second order Kinetics, and our results are also favoring the 2nd order kinetic model [49,50,51].

3.3.4. Regeneration Efficiency of Fe3O4@mZrO2-Re (Re = Y/La/Ce)

Regeneration of materials is one the most important indicator to check the worth of any adsorbent, and the adsorbents with good regeneration abilities are very vital for the cost effective and environmental friendly adsorption process [52]. The regeneration experiment was performed at room temperature where 0.1 g of adsorbent was put in a 250 mL conical flask, and added 100 mL of As3+ and As5+ solutions with a concentration of 100 mg/L, and adjusted the pH to 5. Then placed the conical flask in a constant temperature oscillator for 3 h oscillated at 200 rpm. After filtration, determined the ion concentration in the filtrate to calculate the adsorption capacity. After every cycle NaOH solution was used as an elution agent, then washed with deionized water and dried in oven, and repeated the experiment for seven cycles. The regeneration experiment results of the adsorbent Fe3O4@mZrO2-Re (Re = Y/La/Ce) at different levels are shown in Figure 13.
In the first regeneration cycle, the regeneration efficiencies of both As3+ and As5+ were 93.45% and 95.75%, respectively. It can be seen that the regeneration performance of Fe3O4@mZrO2-Re (Re = Y/La/Ce) for As(III) and As(V) slightly decreases with the increase of regeneration cycles, but the adsorption efficiency of the adsorbent in all the later regeneration experiments tended to be stable. In the 7th cycle, the adsorption capacities of both As3+ and As5+ were 75.15% and 77.59%, respectively. The amount of adsorption slightly decreased after every cycle, and the possible reason for this loss is the structural uncertainty of the adsorbents, which during regeneration treatment might lose their adsorption sites and the structural order accessibility in the pores [53]. By regenerating the nanomaterials during wastewater treatment, we can make the removal process more viable and economical.

4. Conclusions

The optimum synthesis and doping ratios of the Fe3O4@mZrO2 (Re = Y/La/Ce) were determined by uniform design, surface response methodology, and L9 orthogonal design. Based on TEM, SEM, N2 adsorption desorption test, VSM, and XRD results, the suitable molar ratio of Fe3O4:TMAOH:Zirconium Oxide:Y:La:Ce was determined as 1:12:11:1:0.02:0.08. Prepared materials were characterized for morphology, crystallinity, magnetism and porosity etc. The prepared adsorbent materials were mostly spherical particles, and the outer layer of zirconium is mainly in the form of amorphous zirconia, which has good magnetic separation performance, specific surface area, and pore structure. The magnetic saturation capacity of Fe3O4, and Fe3O4@mZrO2-Re (Re = Y/La/Ce) were 49 emu/g and 19.8 emu/g, respectively. The specific surface area was 263 m2/g, porosity was 0.156 cm3/g, and the surface has rich hydroxyl groups. The optimal initial pH values of the system were 6 (As3+) and 5 (As3+). The results of isothermal curves and fitting equation parameters shown that Langmuir model and Redlich Peterson model fitted well, while Freundlich model has a low fitting degree. As per calculations of the Langmuir model the highest adsorption capacities for As3+ and As5+ ions were recorded as 68.33 mg/g, 84.23 mg/g, respectively. From the fitting curves of the selected kinetic models, the adsorption process is more favors the quasi second order kinetic model, while first kinetic order was on lower side. The material regeneration was very effective, so it is suggested to be used as a potential adsorbent for the removal of arsenic-based pollutants from water.

Author Contributions

E.A. completed the methodology, preparation of materials, characterizations, adsorption experiments, and manuscript writing. Q.F. helped in experiment design, supervision, funding, and expert opinion. H.Y. contributed to the methodology, supervision, material arrangement, and experiment design. J.F. helped in Uniform design, Response Surface Methodology, orthogonal design. S.M. helped in interpretation of data, and proof reading of the manuscript. F.B. helped in adsorption kinetic studies using Origin Lab and revisions. All authors have read and agreed to the published version of the manuscript.

Funding

The financial support for this study was received from the Fundamental Research Funds for the central Universities, China, and the grant # 2017QNB05.

Data Availability Statement

The data is available on reasonable request from the corresponding author or first author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Singh, M.K.; Dwivedi, S.; Yadav, S.S.; Sharma, P.; Khattri, S. Arsenic-induced hepatic toxicity and its attenuation by fruit extract of emblica officinalis (Amla) in mice. Indian J. Clin. Biochem. 2014, 29, 29–37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Jiang, Y.; Zeng, X.; Fan, X.; Chao, S.; Zhu, M.; Cao, H. Levels of arsenic pollution in daily foodstuffs and soils and its associated human health risk in a town in Jiangsu Province, China. Ecotoxicol. Environ. Saf. 2015, 122, 198–204. [Google Scholar] [CrossRef] [PubMed]
  3. Ratnaike, R.N. Acute and chronic arsenic toxicity. Postgrad. Med. J. 2003, 79, 391–396. [Google Scholar] [CrossRef]
  4. Neppolian, B.; Doronila, A.; Ashokkumar, M. Sonochemical oxidation of arsenic(III) to arsenic(V) using potassium peroxydisulfate as an oxidizing agent. Water Res. 2010, 44, 3687–3695. [Google Scholar] [CrossRef] [PubMed]
  5. Sun, L.; Lu, M.; Li, Q.; Jiang, H.; Yin, S. Research progress of arsenic removal from wastewater. IOP Conf. Ser. Earth Environ. Sci. 2019, 218, 012142. [Google Scholar] [CrossRef]
  6. Maliyekkal, S.M.; Shukla, S.; Philip, L.; Nambi, I.M. Enhanced fluoride removal from drinking water by magnesia-amended activated alumina granules. Chem. Eng. J. 2008, 140, 183–192. [Google Scholar] [CrossRef]
  7. Sumathi, T.; Alagumuthu, G. Adsorption studies for arsenic removal using activated Moringa oleifera. Int. J. Chem. Eng. 2014, 2014, 1–7. [Google Scholar] [CrossRef] [Green Version]
  8. Zeng, H.; Yu, Y.; Wang, F.; Zhang, J.; Li, D. Arsenic(V) removal by granular adsorbents made from water treatment residuals materials and chitosan. Colloids Surf. A Physicochem. Eng. Asp. 2020, 585, 124036. [Google Scholar] [CrossRef]
  9. Luo, X.; Wang, C.; Wang, L.; Deng, F.; Luo, S.; Tu, X.; Au, C. Nanocomposites of graphene oxide-hydrated zirconium oxide for simultaneous removal of As(III) and As(V) from water. Chem. Eng. J. 2013, 220, 98–106. [Google Scholar] [CrossRef]
  10. Fan, Q.; Cui, X.; Guo, H.; Xu, Y.; Zhang, G.; Peng, B. Application of rare earth-doped nanoparticles in biological imaging and tumor treatment. J. Biomater. Appl. 2020, 35, 237–263. [Google Scholar] [CrossRef]
  11. Fang, K.T.; Lin, D.K.J.; Winker, P.; Zhang, Y. Uniform design: Theory and application. Technometrics 2000, 42, 237–248. [Google Scholar] [CrossRef]
  12. Song, J.; Song, Z.; Sun, R. Study of uniform experiment design method applying to civil engineering. Procedia Eng. 2012, 31, 739–745. [Google Scholar] [CrossRef] [Green Version]
  13. Liang, Y.Z.; Fang, K.T.; Xu, Q.S. Uniform design and its applications in chemistry and chemical engineering. Chemom. Intell. Lab. Syst. 2001, 58, 43–57. [Google Scholar] [CrossRef]
  14. Kleijnen, J.P.C.; Den Hertog, D.; Angün, E. Response surface methodology’s steepest ascent and step size revisited. Eur. J. Oper. Res. 2004, 159, 121–131. [Google Scholar] [CrossRef] [Green Version]
  15. Yang, H.; Fan, J.; Tian, H.; Wang, X.; Fu, W.; Alam, E. Synthesis of imprinted amino-functionalized mesoporous silica and their selective adsorption performance of Pb2+, Cu2+, and Zn2+. J. Sol-Gel Sci. Technol. 2019, 90, 465–477. [Google Scholar] [CrossRef]
  16. Alam, E.; Feng, Q.; Yang, H.; Fan, J.; Mumtaz, S. Synthesis of magnetic core-shell amino adsorbent by using uniform design and response surface analysis (RSM) and its application for the removal of Cu2+, Zn2+, and Pb2+. Environ. Sci. Pollut. Res. 2021. [Google Scholar] [CrossRef]
  17. Kim, T.Y.; Lee, S.H. Combustion and Emission Characteristics of Wood Pyrolysis Oil-Butanol Blended Fuels in a Di Diesel Engine. Int. J. Automot. Technol. 2012, 13, 293–300. [Google Scholar] [CrossRef]
  18. Peasura, P. Application of Response Surface Methodology for Modeling of Postweld Heat Treatment Process in a Pressure Vessel Steel ASTM A516 Grade 70. Sci. World J. 2015, 2015, 318475. [Google Scholar] [CrossRef] [Green Version]
  19. Wu, H. Application of orthogonal experimental design for the automatic software testing. Appl. Mech. Mater. 2013, 347–350, 812–818. [Google Scholar] [CrossRef]
  20. Tada, K.; Nagaya, Y.; Kunieda, S.; Suyama, K.; Fukahori, T. Development and verification of a new nuclear data processing system FRENDY. J. Nucl. Sci. Technol. 2017, 54, 806–817. [Google Scholar] [CrossRef]
  21. Luo, F.; Feng, X.; Jiang, X.; Zhou, A.; Xie, P.; Wang, Z.; Tao, T.; Wan, J. Lanthanum molybdate/magnetite for selective phosphate removal from wastewater: Characterization, performance, and sorption mechanisms. Environ. Sci. Pollut. Res. 2021, 28, 4342–4351. [Google Scholar] [CrossRef]
  22. Kane, S.N.; Mishra, A.; Dutta, A.K. Preface: International Conference on Recent Trends in Physics (ICRTP 2016). J. Phys. Conf. Ser. 2016, 755, 1–7. [Google Scholar] [CrossRef]
  23. Chevallier, S.; Colonna, P.; Buléon, A.; Della Valle, G. Physicochemical behaviors of sugars, lipids, and gluten in short dough and biscuit. J. Agric. Food Chem. 2000, 48, 1322–1326. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, C.; Shaw, L.L. On synthesis of Fe2SiO4/SiO2 and Fe2O3/SiO2 composites through sol–gel and solid-state reactions. J. Sol-Gel Sci. Technol. 2014, 72, 602–614. [Google Scholar] [CrossRef]
  25. Liang, X.; Huan, W.; Jia, X.; Ding, H.; Yang, Y.; Liu, X. Preparation, characterization and magnetic properties of nano-Fe3O4 and cobalt-doped nano-Fe3O4 by co-precipitation method. Nano 2010, 5, 203–214. [Google Scholar] [CrossRef]
  26. Zhang, J.; Zhai, S.; Li, S.; Xiao, Z.; Song, Y.; An, Q.; Tian, G. Pb(II) removal of Fe3O4@SiO2-NH2 core-shell nanomaterials prepared via a controllable sol-gel process. Chem. Eng. J. 2013, 215–216, 461–471. [Google Scholar] [CrossRef]
  27. Zheng, M.; Zhang, H.; Gong, X.; Xu, R.; Xiao, Y.; Dong, H.; Liu, X. A simple additive-free approach for the synthesis of uniform manganese monoxide nanorods with large specific surface area. Nanoscale Res. Lett. 2013, 2013, 3–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Nagy, B.; Mânzatu, C.; Măicăneanu, A.; Indolean, C.; Barbu-Tudoran, L.; Majdik, C. Linear and nonlinear regression analysis for heavy metals removal using Agaricus bisporus macrofungus. Arab. J. Chem. 2017, 10, S3569–S3579. [Google Scholar] [CrossRef] [Green Version]
  29. Shehzad, K.; Ahmad, M.; Xie, C.; Zhan, D.; Wang, W.; Li, Z.; Xu, W.; Liu, J. Mesoporous zirconia nanostructures (MZN) for adsorption of As(III) and As(V) from aqueous solutions. J. Hazard. Mater. 2019, 373, 75–84. [Google Scholar] [CrossRef]
  30. Xu, W.; Wang, J.; Wang, L.; Sheng, G.; Liu, J.; Yu, H.; Huang, X.J. Enhanced arsenic removal from water by hierarchically porous CeO2-ZrO2 nanospheres: Role of surface- and structure-dependent properties. J. Hazard. Mater. 2013, 260, 498–507. [Google Scholar] [CrossRef]
  31. Zhang, G.; Qu, J.; Liu, H.; Liu, R.; Wu, R. Preparation and evaluation of a novel Fe-Mn binary oxide adsorbent for effective arsenite removal. Water Res. 2007, 41, 1921–1928. [Google Scholar] [CrossRef]
  32. Khan, T.A.; Chaudhry, S.A.; Ali, I. Equilibrium uptake, isotherm and kinetic studies of Cd(II) adsorption onto iron oxide activated red mud from aqueous solution. J. Mol. Liq. 2015, 202, 165–175. [Google Scholar] [CrossRef]
  33. Ayawei, N.; Ekubo, A.T.; Wankasi, D.; Dikio, E.D. Adsorption of congo red by Ni/Al-CO3: Equilibrium, thermodynamic and kinetic studies. Orient. J. Chem. 2015, 31, 1307–1318. [Google Scholar] [CrossRef]
  34. Brouers, F.; Al-Musawi, T.J. On the optimal use of isotherm models for the characterization of biosorption of lead onto algae. J. Mol. Liq. 2015, 212, 46–51. [Google Scholar] [CrossRef]
  35. Lakherwal, D. Adsorption of Heavy Metals: A Review. Int. J. Environ. Res. Dev. 2014, 4, 2249–3131. [Google Scholar] [CrossRef]
  36. Wang, J.; Zheng, S.; Shao, Y.; Liu, J.; Xu, Z.; Zhu, D. Journal of Colloid and Interface Science Amino-functionalized Fe3O4@SiO2 core—Shell magnetic nanomaterial as a novel adsorbent for aqueous heavy metals removal. J. Colloid Interface Sci. 2010, 349, 293–299. [Google Scholar] [CrossRef] [PubMed]
  37. Wu, Z.; Li, W.; Webley, P.A.; Zhao, D. General and controllable synthesis of novel mesoporous magnetic iron oxide@carbon encapsulates for efficient arsenic removal. Adv. Mater. 2012, 24, 485–491. [Google Scholar] [CrossRef]
  38. Feng, C.; Aldrich, C.; Eksteen, J.J.; Arrigan, D.W.M. Removal of arsenic from alkaline process waters of gold cyanidation by use of γ-Fe2O3@ZrO2 nanosorbents. Hydrometallurgy 2017, 174, 71–77. [Google Scholar] [CrossRef]
  39. Cui, H.; Li, Q.; Gao, S.; Shang, J.K. Strong adsorption of arsenic species by amorphous zirconium oxide nanoparticles. J. Ind. Eng. Chem. 2012, 18, 1418–1427. [Google Scholar] [CrossRef]
  40. Wang, C.; Luo, H.; Zhang, Z.; Wu, Y.; Zhang, J.; Chen, S. Removal of As(III) and As(V) from aqueous solutions using nanoscale zero valent iron-reduced graphite oxide modified composites. J. Hazard. Mater. 2014, 268, 124–131. [Google Scholar] [CrossRef] [PubMed]
  41. Ma, Y.; Zheng, Y.M.; Chen, J.P. A zirconium based nanoparticle for significantly enhanced adsorption of arsenate: Synthesis, characterization and performance. J. Colloid Interface Sci. 2011, 354, 785–792. [Google Scholar] [CrossRef]
  42. Deng, S.; Li, Z.; Huang, J.; Yu, G. Preparation, characterization and application of a Ce-Ti oxide adsorbent for enhanced removal of arsenate from water. J. Hazard. Mater. 2010, 179, 1014–1021. [Google Scholar] [CrossRef]
  43. Yu, Y.; Zhang, C.; Yang, L.; Paul Chen, J. Cerium oxide modified activated carbon as an efficient and effective adsorbent for rapid uptake of arsenate and arsenite: Material development and study of performance and mechanisms. Chem. Eng. J. 2017, 315, 630–638. [Google Scholar] [CrossRef]
  44. Zhang, L.; Zhu, T.; Liu, X.; Zhang, W. Simultaneous oxidation and adsorption of As(III) from water by cerium modified chitosan ultrafine nanobiosorbent. J. Hazard. Mater. 2016, 308, 1–10. [Google Scholar] [CrossRef]
  45. Sulaymon, A.H.; Mohammed, T.J.; Al-najar, J. Equilibrium and kinetics Studies of Adsorption of Heavy Metals onto Activated Carbon. Can. J. Chem. Eng. Technol. 2012, 3, 86–92. [Google Scholar]
  46. Qiu, H.; Lv, L.; Pan, B.C.; Zhang, Q.J.; Zhang, W.M.; Zhang, Q.X. Critical review in adsorption kinetic models. J. Zhejiang Univ. Sci. A 2009, 10, 716–724. [Google Scholar] [CrossRef]
  47. Pakade, V.; Cukrowska, E.; Darkwa, J.; Torto, N.; Chimuka, L. Selective removal of chromium (VI) from sulphates and other metal anions using an ion-imprinted polymer. Water SA 2011, 37, 529–538. [Google Scholar] [CrossRef] [Green Version]
  48. Alam, M.A.; Shaikh, W.A.; Alam, M.O.; Bhattacharya, T.; Chakraborty, S.; Show, B.; Saha, I. Adsorption of As (III) and As (V) from aqueous solution by modified Cassia fistula (golden shower) biochar. Appl. Water Sci. 2018, 8, 1–14. [Google Scholar] [CrossRef] [Green Version]
  49. Zeng, H.; Zhai, L.; Qiao, T.; Yu, Y.; Zhang, J.; Li, D. Efficient removal of As(V) from aqueous media by magnetic nanoparticles prepared with Iron-containing water treatment residuals. Sci. Rep. 2020, 10, 9335. [Google Scholar] [CrossRef] [PubMed]
  50. Zeng, H.; Qiao, T.; Zhao, Y.; Yu, Y.; Zhang, J.; Li, D. Characterization and arsenic adsorption behaviors of water treatment residuals from waterworks for iron and manganese removal. Int. J. Environ. Res. Public Health 2019, 16, 4912. [Google Scholar] [CrossRef] [Green Version]
  51. Lin, S.; Jin, J.; Sun, S.; Yu, J. Removal of arsenic contaminants using a novel porous nanoadsorbent with superior magnetic recovery. Chem. Eng. Sci. X 2020, 8, 100069. [Google Scholar] [CrossRef]
  52. Huang, X.; Yang, J.; Wang, J.; Bi, J.; Xie, C.; Hao, H. Design and synthesis of core–shell Fe3O4@PTMT composite magnetic microspheres for adsorption of heavy metals from high salinity wastewater. Chemosphere 2018, 206, 513–521. [Google Scholar] [CrossRef] [PubMed]
  53. Liu, A.M.; Hidajat, K.; Kawi, S.; Zhao, D.Y. A new class of hybrid mesaporous materials with functionalized organic monolayers for selective adsorption of heavy metal ions. Chem. Commun. 2000, 15, 1145–1146. [Google Scholar] [CrossRef]
Figure 1. Relationship between Y and Y‘ values.
Figure 1. Relationship between Y and Y‘ values.
Nanomaterials 11 02177 g001
Figure 2. Response surface and contour of As5+ removal for Factor 1 & 3 interaction.
Figure 2. Response surface and contour of As5+ removal for Factor 1 & 3 interaction.
Nanomaterials 11 02177 g002
Figure 3. Response surface and contour of As5+ removal under the interaction of Factor 2 and 3.
Figure 3. Response surface and contour of As5+ removal under the interaction of Factor 2 and 3.
Nanomaterials 11 02177 g003
Figure 4. TEM of adsorbents, (a) uncoated Fe3O4 and (b) Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Figure 4. TEM of adsorbents, (a) uncoated Fe3O4 and (b) Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Nanomaterials 11 02177 g004
Figure 5. VSM Spectra of Fe3O4 and Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Figure 5. VSM Spectra of Fe3O4 and Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Nanomaterials 11 02177 g005
Figure 6. Energy Dispersive spectroscopy (EDS) patterns for Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Figure 6. Energy Dispersive spectroscopy (EDS) patterns for Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Nanomaterials 11 02177 g006
Figure 7. XRD spectrum of Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Figure 7. XRD spectrum of Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Nanomaterials 11 02177 g007
Figure 8. FTIR spectrum of Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Figure 8. FTIR spectrum of Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Nanomaterials 11 02177 g008
Figure 9. N2 adsorption-desorption Isotherm and Pore size distribution curve of Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Figure 9. N2 adsorption-desorption Isotherm and Pore size distribution curve of Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Nanomaterials 11 02177 g009
Figure 10. Effect of pH on As3+ and As5+ adsorption.
Figure 10. Effect of pH on As3+ and As5+ adsorption.
Nanomaterials 11 02177 g010
Figure 11. Fitting of Fe3O4@mZrO2-Re (Re = Y/La/Ce) for adsorption isotherm of As3+ and As5+.
Figure 11. Fitting of Fe3O4@mZrO2-Re (Re = Y/La/Ce) for adsorption isotherm of As3+ and As5+.
Nanomaterials 11 02177 g011
Figure 12. Quasi-first-order kinetics and quasi-second-order adsorption kinetic curves of As(III) and As(V).
Figure 12. Quasi-first-order kinetics and quasi-second-order adsorption kinetic curves of As(III) and As(V).
Nanomaterials 11 02177 g012aNanomaterials 11 02177 g012b
Figure 13. Regeneration cycles of As(III), As(V).
Figure 13. Regeneration cycles of As(III), As(V).
Nanomaterials 11 02177 g013
Table 1. Table of uniform design U14*(145).
Table 1. Table of uniform design U14*(145).
Independent VariableIIIIIIIVV
Experiment Seq.
1st1471113
2nd2814711
3rd312639
4th4113147
5th555105
6th691263
7th713421
8th82111314
9th963912
10th101010510
11th1114218
12th1239126
13th137184
14th1411842
Table 2. Guidelines for selecting columns of generating vectors in U14*(145).
Table 2. Guidelines for selecting columns of generating vectors in U14*(145).
Number of VariablesColumn NumberDeviation
2I, IV0.0957
3I, II, III0.1455
4I, II, III, V0.2091
Table 3. Dose for selected parameters.
Table 3. Dose for selected parameters.
VariableLevel
1234567
Fe3O4 Dosage (X1)/mol0.00040.000450.00050.000550.00060.000650.0007
TMAOH Dosage(X2)/mol0.0040.00450.0050.00550.0060.00650.007
Zirconium-Butoxide(X3)/mol0.0030.00350.0040.00450.0050.00550.006
Table 4. Arrangements of uniform design and test results.
Table 4. Arrangements of uniform design and test results.
Experiment SequenceCodeActual ValueExperimental ValuesPredicted Values
X1X2X3Fe3O4 (mol) (X1)TMAOH (mol) (X2)Zirconium(IV) Butoxide (mol) (X3)Y (mmol/g)Y′ (mmol/g)
1st1470.00040.00450.00450.63220.6122
2nd28140.00040.00550.0060.66430.6443
3rd31260.000450.00650.0040.52240.5324
4th41130.000450.0040.0060.53560.5556
5th5550.00050.0050.0040.66360.6436
6th69120.00050.0060.00550.65670.6767
7th71340.000550.0070.00350.46430.4543
8th82110.000550.0040.00550.67430.6943
9th9630.00060.0050.00350.59280.5728
10th1010100.00060.0060.0050.64750.6875
11th111420.000650.0070.0030.45530.4353
12th12390.000650.00450.0050.62740.6174
13th13710.00070.00550.0030.54470.5247
14th141180.00070.00650.00450.57830.5913
Table 5. Orthogonal Design arrangement of factors.
Table 5. Orthogonal Design arrangement of factors.
NumberFactorsResults
A (Y)B (La)C (Ce)D (Empty List)(mg/g)
11 (0.5%)1 (0.2%)1 (0.5%)129.52
212 (0.4%)2 (0.8%)230.05
313 (0.8%)3 (1.2%)329.16
42 (1%)12332.15
5223131.32
6231230.89
73 (2%)13229.16
8321327.65
9332128.54
Excellent levelA2B1C2
Excellent combinationA2B1C2
Table 6. Elemental composition of Fe3O4@mZrO2-Re (Re = Y/La/Ce).
Table 6. Elemental composition of Fe3O4@mZrO2-Re (Re = Y/La/Ce).
S/NoElementBefore DopingAfter Doped
Weight %Weight %
1O38.1637.2
2Fe9.0812.1
3Zr52.7648.1
4Ce0.1<1.02
5Y0.1<1.08
6La0.1<0.45
7Si0.1<0.1<
10099.95
Table 7. Adsorption Isothermal data of Fe3O4@mZrO2-Re (Re = Y/La/Ce) for As3+ and As5+.
Table 7. Adsorption Isothermal data of Fe3O4@mZrO2-Re (Re = Y/La/Ce) for As3+ and As5+.
ModelParametersAs(III)As(V)
Temperature (K)
298308318298308318
Langmuirqm (mg/g)50.1754.3168.3359.563.6784.23
KL (L/mg)0.1020.0480.0670.1360.0950.104
RL0.6620.8060.7480.5940.6770.556
R20.9870.9920.9910.9910.9910.994
Freundlich K F17.4220.9918.7425.428.9229.45
n 3.97 4.23 3.22 4.87 5.34 3.92
R20.8640.7810.8410.7530.7930.87
Redlich–Peterson K6.866.697.3610.0910.9610.4
α0.1010.0531.120.1130.0990.081
β1.051.161.121.0771.1081.077
R20.9780.9620.9740.9530.9660.988
Table 8. Comparison of adsorption capacity of Fe3O4@mZrO2-Re with other relevant adsorbents.
Table 8. Comparison of adsorption capacity of Fe3O4@mZrO2-Re with other relevant adsorbents.
S#Adsorbent Adsorption Capacity mg/gpHReference
As(III)As(V)
1Fe3O4@mZrO2-Re (Re = Y/La/Ce)68.3384.235–6This study
2CeO2-ZrO29.227.56.7–7.1[31]
3Mesoporous Zirconia nanostructures (MZN)105.03110.295.0–9.0[29]
4Magnetic Fe3O4@Carbon Encapsulates17.9024.407.0[37]
5r-Fe2O3@ZrO262.218.39.0[38]
6Amorphous ZrO2 Nanomaterials8332.47.0[39]
7NZVI-RGO nanoparticles35.829.047.0[40]
8Ce-Al Nanostructured mixed oxiden.a215.0[22]
9Zirconium based nanoparticlesn.a358.0[41]
10Ce(III) doped Titanium-Oxide55.344.96.0[42]
11CeO2 modified Activated carbon36.7743.605.0[43]
12Ce modified chitosan ultrafine nanobiosorbent57.5n.a8[44]
Table 9. Quasi-first-order and quasi-second-order adsorption kinetic fitting data for As(III) and As(V).
Table 9. Quasi-first-order and quasi-second-order adsorption kinetic fitting data for As(III) and As(V).
Metal Initial Concentration
(mg/L)
Quasi First Order Dynamics ModelQuasi Second-Order Dynamics Model
R2qe (cal) (mg/g)K1 min−1R2qe (cal) (mg/g)K2 g mg−1 min−1
As3+10.9879.80.0060.9969.80.0013
20.97223.50.0180.99523.50.0015
50.93443.50.0140.997440.0023
As5+10.9219.50.0210.99311.50.0015
50.95446.50.0220.99150.750.0030
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alam, E.; Feng, Q.; Yang, H.; Fan, J.; Mumtaz, S.; Begum, F. Synthesis of Fe3O4@mZrO2-Re (Re = Y/La/Ce) by Using Uniform Design, Surface Response Methodology, and Orthogonal Design & Its Application for As3+ and As5+ Removal. Nanomaterials 2021, 11, 2177. https://doi.org/10.3390/nano11092177

AMA Style

Alam E, Feng Q, Yang H, Fan J, Mumtaz S, Begum F. Synthesis of Fe3O4@mZrO2-Re (Re = Y/La/Ce) by Using Uniform Design, Surface Response Methodology, and Orthogonal Design & Its Application for As3+ and As5+ Removal. Nanomaterials. 2021; 11(9):2177. https://doi.org/10.3390/nano11092177

Chicago/Turabian Style

Alam, Easar, Qiyan Feng, Hong Yang, Jiaxi Fan, Sameena Mumtaz, and Farida Begum. 2021. "Synthesis of Fe3O4@mZrO2-Re (Re = Y/La/Ce) by Using Uniform Design, Surface Response Methodology, and Orthogonal Design & Its Application for As3+ and As5+ Removal" Nanomaterials 11, no. 9: 2177. https://doi.org/10.3390/nano11092177

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