Vous êtes sur la page 1sur 41

Accepted Manuscript

Title: Optimization of lead removal from aqueous solution


using goethite/chitosan nanocomposite by response surface
methodology

Author: Safoora Rahimi Rozita M. Moattari Laleh Rajabi Ali


Ashraf Derakhshan

PII: S0927-7757(15)30131-X
DOI: http://dx.doi.org/doi:10.1016/j.colsurfa.2015.07.063
Reference: COLSUA 20089

To appear in: Colloids and Surfaces A: Physicochem. Eng. Aspects

Received date: 27-5-2015


Revised date: 23-7-2015
Accepted date: 29-7-2015

Please cite this article as: Safoora Rahimi, Rozita M.Moattari, Laleh Rajabi,
Ali Ashraf Derakhshan, Optimization of lead removal from aqueous solution
using goethite/chitosan nanocomposite by response surface methodology,
Colloids and Surfaces A: Physicochemical and Engineering Aspects
http://dx.doi.org/10.1016/j.colsurfa.2015.07.063

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Optimization of Lead Removal from Aqueous Solution Using

Goethite/Chitosan Nanocomposite by Response Surface Methodology

Safoora Rahimi1, Rozita. M. Moattari1, Laleh Rajabi*1, 2, Ali Ashraf Derakhshan3


1. Polymer Research Lab., Department of Chemical Engineering, College of Engineering, Razi University,
Kermanshah, Iran.
2. Department of Chemistry, University of Victoria (UVic), Victoria, British Columbia, Canada.
3. Young Researchers and Elite Club, Kermanshah Branch, Islamic Azad University, Kermanshah, Iran.

*Corresponding author: Laleh Rajabi


Tel: +1 778 440 3144, +1 250 589 2031
Fax: +98 831 4274542
Email: laleh.rajabii@gmail.com lrajabi@uvic.ca

Graphical abstract

1
Highlights:

 Goethite nanoparticles were used as reinforcement agent in nanocomposite structure.

 Goethite/chitosan nanocomposites were used as sorbents for lead removal.

 Box-bhenken experimental design was used to provide adsorption model.

 The removal efficiency of nanocomposite was greater than pure chitosan film.

Abstract

This work investigates goethite/chitosan nanocomposites for their use in lead removal from

aqueous solutions. Goethite nanoparticles were synthesized and characterized by FTIR, DLS and

SEM. Goethite nanoparticles were identified as nanospheres with the average diameter of 10-60

nm. The optimum conditions were determined using response surface methodology (RSM) based

on three-variable-three-level Box–Behnken design (BBD). The effects of three variables, i.e.

initial solution pH, adsorbent mass and initial concentration of Pb (II) ions on the removal

efficiency for Pb (II) ions were evaluated. The optimal conditions for the lead removal were

found to be 6, 0.05 g and 74.4 mg/L, for the initial solution pH, adsorbent mass and the initial

concentration of Pb (II) ions, respectively. Under these conditions, maximum lead removal

efficiency was obtained to 98.26% that was in respectable agreement with the model (97.19%).

The modified quadratic model exhibited excellent stability for Pb(II) adsorption by

goethite/chitosan nanocomposite. The results of adsorption study by goethite/chitosan

nanocomposite revealed that Pb(II) uptake was enhanced by chitosan film using goethite

nanoparticles.

Keywords: Adsorption, Goethite, Chitosan, Nanocomposite, Nanoparticle, Box–Behnken design

1. Introduction

2
Nowadays, environmental pollution which is a result of rapid technological development is a

serious apprehension for ecosystem. Some pollutants like heavy metals rarely disappear; they are

harmful to humans, animals, and other living creatures. Lead is known as one of the most toxic

heavy metals which is typically resulted from the industrial wastes of the lead mining, lead smelting,

battery manufacturing, printing, pigments, fossil fuels, photographic materials, explosive

manufacturing, rubber production, etc. . High-level lead exposure can adversely damage the brain

and kidneys and even cause death. Because of the high toxicity of lead ion, it’s imperative to remove

the lead ion from waste water before discharging it into the environment. For this purpose, A variety

of methods have been employed to remove lead ions from industrial wastewaters, such as solvent

extraction, precipitation and coagulation, biosorption, membrane filtration, chemical absorption, low

energy reverse osmosis, adsorption and so on [1-3]. Among these mentioned methods, adsorption is

widely used due to its high removal efficiency, easy handling, high selectivity and lower operating

cost even at very low concentrations of lead. To date, different types of adsorbents such as

zeolites, metallic oxides, activated carbon, ion exchange resins, polymeric adsorbents and

different biosorbents have been employed for lead removal from waste water. Low adsorption

capacity of these adsorbents usually restricts their large-scale application in water treatment.

Therefore, there is a crucial need for new adsorbents with characteristics such as high adsorption

capacity, easy fixation, and separation from water [1, 3].

Polysaccharide are known as functional and biocompatible materials with a great capability to be

crosslinked with nano and micro structures *Gao. Chitosan, a natural polysaccharide-based

polymer obtained from chitin, is popular in various applications due to its nontoxicity and

biodegradability [4]. It is well known as a low cost adsorbent for heavy metal removal with a

high adsorption capacity consists a large number of functional groups such as amino (NH2) and

hydroxyl (OH) groups [5]. Nowadays, preparation of chitosan nanocomposite with better

3
mechanical and chemical properties from that of the pure chitosan film have been the focus of

attention of several research groups [6, 7]. Activated clay [8], poly vinyl alcohol, poly vinyl

chloride, kaolinite [9], perlite [10], glyoxal, formaldehyde, glutaraldehyde, epichlorohydrin, oil

palm ash [11], ethylene glycon diglycidyl ether and isocyanates [12] montmorillonite [13] and

bentonite [14] have been added as cross-linking agents to improve adsorption properties of

chitosan nanocomposites. Tao and his coworkers reported the use of TiO2 in forming

chitosan/TiO2 hybrid film for the removal of pb(II) from aqueous solution. They used Box–

Behnken model and indicated that the reaction parameter optimization using response surface

method is scientific and valid [15]. RSM is essentially a particular collection of mathematical

and statistical techniques for designing experiments, building models, evaluating the effects of

variables, and optimizing process. Its greatest advantage is reducing the number of experimental

trials required to evaluate numerous parameters and their interactions. This methodology can be

used in developing suitable treatment technology considering the effects of operational

conditions on the removal process. In recent years, RSM has been applied to optimize and assess

interactive effects of independent parameters in various chemical and biochemical procedures

[16-18] .

This paper reports on chitosan, containing new functional groups in order to increase the density

of adsorption sites. Goethite nanoparticles were synthesized and applied as the reinforcing agent

to improve the adsorption properties of the chitosan film. Three operating parameters including

initial solution pH (pH), nanoparticle dose (Cn) and initial concentration of Pb(II) ions (C0) were

investigated. The Box–Behnken model was used to statistically design the experiments to assess

and optimize adsorption process, using Design Expert 7 software.

4
2. Experimental

2.1. Reagents and materials

Chitosan powder with low molecular weight was purchased from Sigma-Aldrich (USA). NaOH,

HNO3, acetic acid, ethanol and lead stock solution [NIST Pb(NO3)2 in HNO3 1000 ppm] were

purchased from Merck, Germany. All other reagents used, were of analytical grade and

purchased from Merck, Germany. All the solutions were prepared with deionized water.

2.2. Preparation of goethite/chitosan nanocomposites

Goethite was synthesized from the reaction of Fe(NO3)3·9H2O and KOH [19]. The obtained

suspension was sonicated for 30 minutes at room temperature and then placed in the oven for 70

minutes at 100°C and centrifuged. Later, the modified goethite/chitosan adsorptive

nanocomposites were prepared by solution-casting method. Three goethite/chitosan solutions,

each containing various amounts of goethite were prepared in the following manner. As

described in detail in Table 1, A certain amount of goethite and 0.2 g of chitosan powder were

dissolved in 1% (v/v) of acetic acid. The resulting suspension was bath-sonicated for 10 min to

guarantee complete dissolution of the solutes.

Table 1

The resulting solutions were incubated at room temperature under 180 rpm for 24 h and placed in

the pre-heated oven at 50°C for 48 h to help releasing the entrapped air bubbles. The obtained

solutions were casted on glass plates using a casting knife and dried at room temperature. The

films were immersed into alkaline solution (1M NaOH) at ambient temperature for 1 h in order

to minimize the solubility of the films in water and also neutralize the excess acid. The casted

films were thoroughly washed with deionized water and dried at room temperature for 24 hours.

5
Pure chitosan films were also prepared using the same procedure. The choice of goethite for

preparation of nanocomposite against lepidocrocite is discussed later in the results and discussion

section.

2.3. Characterization

Goethite/chitosan nanocomposites were characterized using FTIR and SEM. Fourier-transform

infrared (Bruker alpha, German) spectra were recorded between 400 and 4000 cm-1 with KBr

pellets at room temperature. SEM images of all samples were taken, using Philips XL-30S FEG

and LEQ 1450 VP. Dynamic light scattering analysis (DLS, Malvern Instruments, UK) was

carried out to determine size distribution and average particle size of nanoparticles. Samples for

DLS analysis were prepared as 5% W/V (nanoparticle/water).

2.4. Adsorption experiments

Dilution of the standard lead solution (1000 mg L-1) with deionized water provided required

initial solutions with appropriate concentration. A known amount of the adsorbent (0.02 g) was

added to 25 ml of the required concentration of Pb (NO3)2 solutions and transferred to a 100 ml

Erlenmeyer flask with a constant agitation rate of 180 rpm at ambient temperature. The pH of the

solutions was adjusted to the required value through the addition of HNO3 (1 M) and NaOH (1

M) solutions. The samples were then filtered to separate the nanocomposite from aqueous

solutions. Supernatants were analyzed using an atomic absorption spectroscopy (SHIMADZO

AA-6300). Lead removal (%) by nanocomposite was determined according to Eq. (1):

6
Where R% is removal percentage and q t is the amount of lead uptake by the adsorbent in mgg−1.

C0 and Ct are the initial and final metal ion concentrations in mgL−1 in the solution, respectively.

V and m indicate the volume of solution (L) and weight of the adsorbent (g), respectively.

2.5. Box–Behnken experimental design

As one of the RSM designs, Box–Behnken design is known as a modified central composite

experimental design [20]. It is demonstrated that Box–Behnken design is more efficient and

requires fewer experiments in comparison to other RSM designs. The benefits of Box–Behnken

designs include the fact that they are all spherical designs and require factors to be run at only

three levels. It can be noted that a number of additional experiments as well as time consuming

and laborious laboratory studies will be eliminated by selecting the Box–Behnken experimental

design, instead of complete factorial design [21].

Pb(II) uptake by goethite/chitosan-nanocomposite was investigated using response surface

methodology (RSM) based on three-variable-three-level Box–Behnken design (BBD). The

Design Expert 7 software was used for regression and graphical analysis of the obtained data.

For statistical calculations, the three independent variables were designed as X1, X2 and X3 with

the coded values at three levels: -1, 0 and +1. The effects of three variables on the removal

efficiency for Pb2+ were evaluated.

3. Results and discussion

3.1. Characterization of adsorbents

Fig. 1 shows the FTIR spectra of goethite nanoparticle, pure chitosan film and goethite/chitosan

nanocomposite. For the goethite spectrum, the two index peaks at 3400 and 3150 cm-1 are

7
assigned to the H-O-H vibration which are related to non-stoichiometric hydroxyl units (excess

water) in the goethite structure. The band at 1634 cm−1 was assigned to the water bending

vibration. Strong absorption peaks at 890 cm−1 and 790 cm−1 caused by the in-plane bending of

surface hydroxyl of Fe-OH-Fe, were similar to the one reported in literature [22, 23]. The main

index peaks of goethite nanoparticles in the spectrum of the gothite/chitpsan nanocomposite are

highlited in the Fig. 1, clearly indicating the involvement of goethite nanoparticles in

nanocomposite structure.

Fig. 1

The average particle size of goethite nanoparticles had been measured to be 145.6 nm through

DLS analysis, smaller than what was reported by Roze and his co-workers [24].

Fig. 2 shows SEM micrographs of goethite nanoparticles. The goethite nanoparticles are

spherical with diameter of 10-70 nm.

Fig. 2

Fig. 3 shows the SEM micrographs from the surface of the goethite/chitosan nanocomposite and

also the three dimensional structure of it. Goethite nanoparticles are dispersed in chitosan

polymeric matrix and the size of particles, including the aggregated ones is given in the picture.

Fig. 3

Fig. 4 presents the cross sectional SEM micrograph of goethite/chitosan nanocomposite.

Thickness of chitosan film was found to be 450-470 µm, approximately.

Fig. 4

Fig. 5 shows the the proposed schematic illustration of goethite/chitosan nanocomposite

structure. Goethite nanoparticles act as cross-linking agents to link the biopolymer chitosan

chains through the formation of both strong primary covalent bonds as well as hydrogen

8
bondings. As a matter of fact, goethite nanoparticles act as both reinforcing agent and adsorbent

for the chitosan polymer matrix.

Fig. 5

3.2. RSM approach and statistical analysis

Response surface methodology (RSM) is a combination of mathematical and statistical

techniques that are useful for modelling and analysis of problems in which output or response is

influenced by several input variables and the objective is to find the correlation between the

response and the variables investigated [25]. A polynomial regression equation was developed

by using Box–Behnken design to analyze the factor interactions. The complete design matrix

together with observed and predicted experimental response values are given in Table 2. Three

replicate at the center point is used to determine the experimental error.

Table 2

3.3. Determination of the regression model and statistical evaluation

In order to fit an empirical second-order polynomial, model response function (Y) for predicting

the optimal point was constructed in the form Eq. (2):

Where Y is a response variable of removal efficiency (%); b 0 is the constant coefficient, bi are

the regression coefficients for linear effects, bii and bij are the square and interaction effects,

respectively. Xi and Xj are the coded experimental levels of the variables and k is the number of

9
the independent variables . The software Design Expert 7 and Minitab 16, were used for the

experimental design, determination of the coefficients, the data analysis and the graph plotting.

By comparing the experimental and predicted values the reliability of the model and the

credibility of the statistical evaluations were determined. The effects of process variables

including initial pH, nanoparticle dose, and initial lead concentration on the lead removal

efficiency were investigated using RSM according to BBD. Different response terms such as

linear, interactive, quadratic and cubic models were used to correlate the experimental data and

to obtain the regression equation. To decide about the competence of the obtained models to

describe lead removal by goethite/chitosan nanocomposite, three different tests (sequential

model sum of squares, lack of fit tests, and model summary statistics) were carried out in the

present study and the results are presented in Table 3.

Table 3

From Table 3, it is evident that quadratic model is the most suitable for the removal of Pb(II) by

the nanocomposite. The competence, significance and compatibility of the model was further

check through analysis of variance (ANOVA).

The ANOVA for the quadratic model for lead inos adsorption onto nanocomposite is tabulated in

Table 4. All terms in the regression models were not equally important. The significance of each

coefficient was determined by F-value and p-values, which are listed below. In general, the

larger the magnitude of the F-value and the smaller the p-value, the more significant is the

corresponding

coefficient term. The Model F-value of 36.40 implied its significance. In this case initial pH

(X1), nanoparticle dose (X2), initial concentration of Pb 2+ ions (X3) and interactions X1X3, X12,

10
X22 and X32 were significant model terms. The p-values ≥0.050 indicated the model terms that

were not significant. So it is better model reduction was employed to improve model; for this

purpose X1X2 and X2X3 were ruled out. It was observed that, model F-value reduced, so X1X2

was reconsidered. With this improvement, model F-value was 47.15 (Table 5). The correlation

coefficients R2 and R2adj were computed to check the adequacy of the model. In statistical

modeling, by removing a repressor variable, the coefficient of determination decreases and a

large value of R2 does not necessarily imply that the regression model is a good one. Hence, R2adj

is preferred to be used to determine the fit of a regression model, as it does not always increase

when variables are added [26, 27]. In the current work, the high value of R2 (0.984)

demonstrated a high dependence and correlation between the observed and the predicted values

of response. The value of R 2adj (0.963) indicated that the total variation of about 96% for lead

removal was attributed to the independent variables and only about 4% of the total variation

cannot be explained by the model. Eq. (3) shows the response functions with the determined

coefficients for Pb(II) removal; the initial pH of solution (X1), nanoparticle dose (X2) and initial

concentration (X3) are represented in terms of coded factors (-1, 0 and +1).

Y = 80.80+ 10.32X1 + 2.67X2 - 6.59X3 - 2.26X1X2 + 3.29X1X3 – 7.89 X22 - 3.23X32

(3

It can be seen from the coefficients in Eq. (3) that removal efficiency increases with the pH (X1)

and Cn (X2) and decreases with C0 (X3). Initial pH (X1) has a more profound effect on lead

adsorption as compared to nanoparticle dose (X2) and initial concentration (X3), which is in

agreement with contribution percentages presented in Fig. 6.

Table 4

11
Table 5

Fig. 6

3.4. Diagnostic analysis of fitted model adequacy

The statistical analysis above verified the appropriate fit of model to the observed lead removal

efficiency. However, these values do not imply the adequacy of the model for its intended

application. This would require a basic diagnostic check of model adequacy. Generally, observed

data can be expressed as follow:

Observed data value = true model + random (4)

Observed data value = fitted model + residual (5)

Comparison of Equations (4) and (5) suggest that the fitted model is close the true model when

the residuals are close to random errors. Random errors are defined as a sequence of independent

and normally distributed observations. The randomness and normality diagnosis of the residuals

from the fitted model and observed adsorption data form the basis for judging the

appropriateness of the fitted model [28]. Fig. 7 shows histogram error plot indicating error

density is centralized at zero.

Fig. 7

The fitted quality of Eq. (3) was also expressed by comparing lead removal efficiency between

experimental and predicted values, as shown in Fig. 8. The better the fit of the model, the smaller

the values of residuals is, more to the point, residuals should be normally distributed [29]. It is

clear from Fig. 8 that the predicted values are quite close to the actual experiment, thus

12
confirming that the regression model exhibits excellent stability for Pb(II) adsorption on

goethite/chitosan nanocomposite. Therefore, it can be concluded that the response surface model

developed in this study (Eq. (3)) was considered to be satisfactory for the prediction of Pb(II)

adsorption system.

Fig. 8
3.5. Comparative effects of media components on Pb (II) removal efficiency

The perturbation plot was used to compare the effects of all the parameters at a point in the

design

space on the response (Fig. 9). A sharp slope for solution pH shows that the response of Pb(II)

removal efficiency was very sensitive to this parameter. The nearly flat curves for nanosorbent

dose indicated that removal efficiency was insensitive to this factor as compared to the solution

pH. Furthermore, perturbation plot for initial concentration (C) shows that this factor can affect

adsorption process considerably; although this is not as influential as pH. It was clear from the

perturbation plot that the most significant factor on the response was the solution pH.

Fig. 9
3.6. Contour plots and response surface analysis

The three-dimensional response surface plots and two-dimensional contour plots are the

graphical representations of the regression equation. These types of plots demonstrate the effects

of two factors on the response at a time. Therefore, in this work 3D response surface plots for the

measured responses were formed based on the model Eq. (3). The relationship between the

dependent and independent variables was further illuminated by constructing contour plots.

Figures 10 and 11 show the 3D response surfaces and the corresponding contour plots as the

functions of two variables at the center level of other variables, respectively.

13
Fig. 10
Fig. 11
3.7. Effects of model components and their interactions on Pb(II) removal efficiency

The solution pH value plays an important role in adsorption process and specifically on the

adsorption capacity of the adsorbent. Solution pH would affect both aqueous chemistry and

surface binding sites of the adsorbents, thus, changing solution pH could modify the surface

charge of an adsorbent. Based on the electron donating nature of the amine (–NH2) and hydroxyl

(–OH) groups in chitosan and the electron accepting nature of Pb2+ ions, it seems that the ion

exchange mechanism could be preferentially considered. In the lower pH region the positively

charged sites dominate and the H+ ions compete with Pb2+ cations for the exchange sites on the

sorbent surface. While the solution pH increases, the number of negatively charged sites

increases, which results in a lower coulombic repulsion of the sorbing metal. In this part of the

the current work, the pH range of 3 - 6 was assayed for lead removal by goethite/chitosan-

nanocomposite. pH 3 was chosen for lower bound due to solubility of chitosan film in water at

pHs lower than 3. When the initial pH of the lead solution with concentration of 150 ppm was

adjusted to values higher than 6.3, lead precipitation (Pb (OH)2) occurred due to the existence of

OH− ions in the adsorption medium. Thus the pH 6.3 was selected as the upper limit bound.

Fig. 10a and b shows the combined effects of pH correspond with nanoparticle dose and initial

concentration, respectively. As seen in fig. 10b at high lead concentrations, altering the solution

pH significantly affected removal efficiency, and this phenomenon is attributed to the influence

of pH in the presence of OH3+ cations. Furthermore, interaction of nanoparticle dose and solution

pH at low Cn values is more considerable in comparison to high Cn values.

14
It can be depicted from the response graphs that the metal removal efficiency is dependent on the

initial metal concentration, in a way that, as the concentration increased the removal efficiency

decreased. At high lead concentrations, metal ions occupy active sites of adsorbent quickly, thus

reducing the number of available adsorption sites. It can be said that, the adsorbent surface is

saturated by lead ions, and this fact prevents the efficient ion adsorption by adsorbent [30].

Figures 10b and 11b well-describe this fact.

3.8. Mechanism of adsorption by PABF/chitosan nanocomposite

Fig. 12 shows the mechanism of Pb(II) adsorption onto the goethite/chitosan nanocomposite.

Both ion exchange and complex adsorptions may occur through the adsorption process. Fig. 12a

and b shows complex formation and ion exchange adsorption mechanisms, respectively.

Fig. 12

3.9. Optimization of variables for removal of Pb(II)

The optimum values of the selected test variables were obtained by solving the Eq. (3) and also

by analyzing the response surface contour plots. The optimum variables were found to be 6 (X1=

1) for initial pH of the solution, 74.4 ppm (X2 = -0.51) for initial concentration of Pb(II) ions,

and 0.05 g (X3 =0.026) for nanoparticle dose with a predicted Pb(II) removal efficiency of about

97.19%. Later, confirming experiment was carried out to assess the predicted result. The

experimental value was 98.26%, which was in well agreement with the predicted value. The

result showed a very good consistency and this indicates that there is a good concurrence

between the predicted and the experimental values.

3.10. Adsorption by pure chitosan film

15
Pb(II) adsorption experiment by pure chitosan film at pH=6, C0= 50, contact time=24 h, with

agitation rate of 180 rpm at ambient temperature was performed. The removal efficiency and the

adsorption capacity were found to be 58.58 and 36.61, respectively; while removal efficiency

and the adsorption capacity in the same condition for goethite/chitosan-nanocomposite (Cn=

0.046) were 98.42 and 61.51, respectively (run 6 of Table 2). A comparison between these data

shows an improvement in Pb(II) uptake by chitosan film using goethite nanoparticles (Fig. 13).

Fig. 13

4. Conclusions

Goethite nanoparticles acted as both nanofiller and adsorbent for the chitosan polymer matrix.

The modified quadratic model exhibited excellent stability for Pb(II) adsorption by

goethite/chitosan nanocomposite. The response surface model developed in this study was

considered to be satisfactory for the prediction of Pb(II) adsorption system R2=0.984. Increasing

the solution pH significantly increased removal efficiency. Pb(II) uptake was enhanced by

chitosan film using goethite nanoparticles. The obtained experimental value in optimum

conditions was 98.26% that was in well agreement with the predicted value (97.19%).

16
References:

[1] M.A. Salam, Coating carbon nanotubes with crystalline manganese dioxide nanoparticles and

their application for lead ions removal from model and real water, Colloids and Surfaces A:

Physicochemical and Engineering Aspects, 419 (2013) 69-79.

[2] L. Jiang, P. Liu, S. Zhao, Magnetic ATP/FA/Poly (AA-co-AM) ternary nanocomposite

microgel as selective adsorbent for removal of heavy metals from wastewater, Colloids and

Surfaces A: Physicochemical and Engineering Aspects, 470 (2015) 31-38.

[3] S. Duan, R. Tang, Z. Xue, X. Zhang, Y. Zhao, W. Zhang, J. Zhang, B. Wang, S. Zeng, D.

Sun, Effective removal of Pb (II) using magnetic Co0.6 Fe2.4O 4 micro-particles as the adsorbent:

Synthesis and study on the kinetic and thermodynamic behaviors for its adsorption, Colloids and

Surfaces A: Physicochemical and Engineering Aspects, (2015).

[4] Y. Yang, Q. He, L. Duan, Y. Cui, J. Li, Assembled alginate/chitosan nanotubes for biological

application, Biomaterials, 28 (2007) 3083-3090.

[5] Q. Zhou, Q. Gao, W. Luo, C. Yan, Z. Ji, P. Duan, One-step synthesis of amino-functionalized

attapulgite clay nanoparticles adsorbent by hydrothermal carbonization of chitosan for removal

of methylene blue from wastewater, Colloids and Surfaces A: Physicochemical and Engineering

Aspects, 470 (2015) 248-257.

[6] M.-S. Chiou, P.-Y. Ho, H.-Y. Li, Adsorption of anionic dyes in acid solutions using

chemically cross-linked chitosan beads, Dyes and Pigments, 60 (2004) 69-84.

[7] Z. Zhu, J. Wang, A. Munir, H.S. Zhou, Direct electrochemistry and electrocatalysis of

horseradish peroxidase immobilized on bamboo shaped carbon nanotubes/chitosan matrix,

Colloids and Surfaces A: Physicochemical and Engineering Aspects, 385 (2011) 91-94.

17
[8] M.-Y. Chang, R.-S. Juang, Adsorption of tannic acid, humic acid, and dyes from water using

the composite of chitosan and activated clay, Journal of Colloid and Interface Science, 278

(2004) 18-25.

[9] H.-Y. Zhu, R. Jiang, L. Xiao, Adsorption of an anionic azo dye by chitosan/kaolin/γ-Fe2O3

composites, Applied Clay Science, 48 (2010) 522-526.

[10] S. Kalyani, J.A. Priya, P.S. Rao, A. Krishnaiah, Removal of copper and nickel from aqueous

solutions using chitosan coated on perlite as biosorbent, Separation science and technology, 40

(2005) 1483-1495.

[11] M. Hasan, A. Ahmad, B. Hameed, Adsorption of reactive dye onto cross-linked chitosan/oil

palm ash composite beads, Chemical Engineering Journal, 136 (2008) 164-172.

[12] G. Crini, P.-M. Badot, Application of chitosan, a natural aminopolysaccharide, for dye

removal from aqueous solutions by adsorption processes using batch studies: A review of recent

literature, Progress in polymer science, 33 (2008) 399-447.

[13] L. Wang, J. Zhang, A. Wang, Removal of methylene blue from aqueous solution using

chitosan-g-poly (acrylic acid)/montmorillonite superadsorbent nanocomposite, Colloids and

Surfaces A: Physicochemical and Engineering Aspects, 322 (2008) 47-53.

[14] W.S.W. Ngah, N.F.M. Ariff, M.A.K.M. Hanafiah, Preparation, characterization, and

environmental application of crosslinked chitosan-coated bentonite for tartrazine adsorption from

aqueous solutions, Water, Air, and Soil Pollution, 206 (2010) 225-236.

[15] Y. Tao, L. Ye, J. Pan, Y. Wang, B. Tang, Removal of Pb (II) from aqueous solution on

chitosan/TiO2 hybrid film, Journal of hazardous materials, 161 (2009) 718-722.

[16] B.K. Körbahti, M. Rauf, Response surface methodology (RSM) analysis of photoinduced

decoloration of toludine blue, Chemical Engineering Journal, 136 (2008) 25-30.

18
[17] S. Sharma, A. Malik, S. Satya, Application of response surface methodology (RSM) for

optimization of nutrient supplementation for Cr (VI) removal by Aspergillus lentulus AML05,

Journal of hazardous materials, 164 (2009) 1198-1204.

[18] M. Amini, H. Younesi, N. Bahramifar, A.A.Z. Lorestani, F. Ghorbani, A. Daneshi, M.

Sharifzadeh, Application of response surface methodology for optimization of lead biosorption

in an aqueous solution by Aspergillus niger, Journal of hazardous materials, 154 (2008) 694-702.

[19] S. Rahimi, R.M. Moattari, L. Rajabi, A.A. Derakhshan, M. Keyhani, Iron oxide/hydroxide

(α, γ-FeOOH) nanoparticles as high potential adsorbents for lead removal from polluted aquatic

media, Journal of Industrial and Engineering Chemistry, (2014).

[20] J. Ahmad, S.R. Mir, K. Kohli, S. Amin, Effect of oil and co-surfactant on the formation of

Solutol HS 15 based colloidal drug carrier by Box–Behnken statistical design, Colloids and

Surfaces A: Physicochemical and Engineering Aspects, 453 (2014) 68-77.

[21] K. Yetilmezsoy, S. Demirel, R.J. Vanderbei, Response surface modeling of Pb (II) removal

from aqueous solution by Pistacia vera L.: Box–Behnken experimental design, Journal of

hazardous materials, 171 (2009) 551-562.

[22] S. Liao, J. Wang, D. Zhu, L. Ren, J. Lu, M. Geng, A. Langdon, Structure and Mn2+

adsorption properties of boron-doped goethite, Applied Clay Science, 38 (2007) 43-50.

[23] H. Ruan, R.L. Frost, J.T. Kloprogge, L. Duong, Infrared spectroscopy of goethite

dehydroxylation. II. Effect of aluminium substitution on the behaviour of hydroxyl units,

Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy, 58 (2002) 479-491.

[24] M.a.M. Cortalezzi, J. Rose, G.F. Wells, J.-Y. Bottero, A.R. Barron, M.R. Wiesner, Ceramic

membranes derived from ferroxane nanoparticles: a new route for the fabrication of iron oxide

ultrafiltration membranes, Journal of membrane science, 227 (2003) 207-217.

19
[25] M. Pradhan, C. Biswas, Modeling and Analysis of process parameters on Surface

Roughness in EDM of AISI D2 tool Steel by RSM Approach, International Journal of

Mathematical, Physical and Engineering Sciences, 3 (2009).

[26] A. Habibi, F. Vahabzadeh, Formaldehyde degradation by Ralstonia eutropha in an

immobilized cell bioreactor, Journal of Environmental Science and Health, Part A, 48 (2013)

1557-1572.

[27] Z. Zhang, H. Zheng, Optimization for decolorization of azo dye acid green 20 by ultrasound

and H2O2 using response surface methodology, Journal of hazardous materials, 172 (2009) 1388-

1393.

[28] K. Zhang, W. Cheung, M. Valix, Roles of physical and chemical properties of activated

carbon in the adsorption of lead ions, Chemosphere, 60 (2005) 1129-1140.

[29] M.A. Islam, V. Sakkas, T.A. Albanis, Application of statistical design of experiment with

desirability function for the removal of organophosphorus pesticide from aqueous solution by

low-cost material, Journal of hazardous materials, 170 (2009) 230-238.

[30] G. Zolfaghari, A. Esmaili-Sari, M. Anbia, H. Younesi, S. Amirmahmoodi, A. Ghafari-

Nazari, Taguchi optimization approach for Pb (II) and Hg (II) removal from aqueous solutions

using modified mesoporous carbon, Journal of hazardous materials, 192 (2011) 1046-1055.

Figures

Figure 1. FTIR spectra of goethite nanoparticle, pure chitosan film and goethite/chitosan

nanocomposite

Figure 2. SEM micrographs of goethite nanoparticles

Figure 3. SEM image of goethite/chitosan nanocomposite surface

20
Figure 4. Cross-sectional SEM image of goethite/chitosan nanocomposite

Figure 5. The proposed structure of goethite /chitosan nanocomposite made from chitosan

biopolymer and goethite nanoparticles

Figure 6. Percent contribution of various parameters on lead removal efficiency by

goethite/chitosan nanocomposite.

Figure 7. Histogram plot of errors, Mean: mean of errors, StDeV: standard deviation, N: number

of experiments

Figure 8. Correlation of observed and predicted lead adsorption

Figure 9. Perturbation plots; (A) initial solution pH, (B) nanoparticle dose and (C) initial lead

concentration

Figure 10. 3-D surface plots for interactive effect of (a) pH and adsorbent dose while initial

concentration was adjusted in 100 mg/L (b) pH and initial concentration while Cn was adjusted

in 0.046 g

Figure 11. Contour plots exhibiting the interactive effects between two independent variables

(other variables were held at their respective center levels); (a) initial pH of solution (pH, X1)

and nanosorbent dose (Cn, X2), (b) initial pH of solution (pH, X1) and initial concentration of

Pb(II) ions (Cn, X3)

Figure 12. Mechanism of Pb (II) adsorption onto the goethite/chitosan nanocomposite, (a)

complex formation mechanism and (b) ion exchange mechanism

Figure 13. Comparison between removal efficiency and capacity of pure chitosan and goethite/

chitosan films (pH=6, C0= 50 and contact time=24 h)


21
Figure. 1

22
Figure. 2

23
Figure. 3

24
Figure. 4

25
Figure. 5

26
Figure. 6

27
Figure. 7

28
Figure. 8

29
Perturbation
99
A

89

C
R%

79
A B

B C
69

59

-1.000 -0.500 0.000 0.500 1.000

Deviation from Reference Point (Coded Units)

Figure. 9

30
Figure. 10

31
Figure. 11

32
Figure. 12

33
Figure. 13

Tables

Table 1. Synthesis details of Goethite/chitosan nanocomposite

Table 2. Complete design matrix with observed and predicted experimental response values

Table 3. Sequential model fitting for the lead adsorption on goethite/chitosan nanocomposite

34
Table 4. ANOVA for Response Surface Full Quadratic Model

Table 5. ANOVA for Response Surface Reduced Quadratic Model

35
Table 1

Total weight (g) 10 10 15

Nanoparticle mass (g) 0.046 0.09 0.003

Chitosan mass (g) 0.2 0.2 0.3

Acid mass (g) 9.754 9.71 14.697

36
Table 2
Run Independent variables Removal
pH, x1 Nanoparticle initial efficiency (%)
dose, x2 concentration, x3
Coded Actual Coded Actual Coded Actual Observed Predicted
(g) (mg/l)
1 -1 3 -1 0.002 0 100 64.94 62.87
2 1 6 -1 0.002 0 100 88.18 88.04
3 -1 3 1 0.09 0 100 72.61 72.74
4 1 6 1 0.09 0 100 86.79 88.86
5 -1 3 0 0.046 -1 50 82.43 82.35
6 1 6 0 0.046 -1 50 98.43 96.42
7 -1 3 0 0.046 1 150 60.58 62.59
8 1 6 0 0.046 1 150 89.72 89.80
9 0 4 -1 0.002 -1 50 71.98 73.60
10 0 4.5 1 0.09 -1 50 78.48 78.95
11 0 4.5 -1 0.002 1 150 59.84 60.42
12 0 4.5 1 0.09 1 150 68.43 65.76
13 0 4.5 0 0.046 0 100 80.97 80.80
14 0 4.5 0 0.046 0 100 80.40 80.80
15 0 4.5 0 0.046 0 100 81.01 80.80

37
Table 3
Sequential model sum of squares
Sum of DOF Mean square F-value P-value > Comme
Source squares F nt
Mean vs. Total 90447.09 1 90447.09 - - -
Linear vs. Mean 1256.98 3 418.99 9.58 0.0021 -
2FI vs. Linear 64.78 3 21.59 0.41 0.7470 -
Quadratic vs. 390.14 3 130.05 24.89 0.0020 Suggest
2FI ed
Cubic vs. 25.90 3 8.63 74.74 0.0132 Aliased
Quadratic
Residual 0.23 2 0.12 - - -
Total 92185.13 15 6145.68 - - -
Lack of Fit Tests
Sum of DO Mean square F-value P-value > Remark
Source squares F F
Linear 480.82 9 53.42 462.53 0.0022 -
2FI 416.04 6 69.34 600.33 0.0017 -
Quadratic 25.90 3 8.63 74.74 0.0132 Suggest
ed
Cubic 0.000 0 - - - Aliased
Pure Error 0.23 2 0.12 - - -
Model Summary Statistics
Std. Dev. R2 R2adj Predicted PRESS Remark
Source R2
Linear 6.61 0.7232 0.6477 0.4243 455.24 -
2FI 7.21 0.7605 0.5809 -0.2037 485.90 -
Quadratic 2.29 0.9850 0.9579 0.7613 269.04 Suggest
ed
Cubic 0.34 0.9999 0.9991 - - Aliased

38
Table 4
Source Sum of DOF Mean F-value p-value > F
squares square
Model 1711.91 9 190.21 36.40 0.0005
X1-pH 852.15 1 852.15 163.07 < 0.0001
X2-C n 57.15 1 57.15 10.94 0.0213
X3-C0 347.68 1 347.68 66.54 0.0004
X1X2 20.48 1 20.48 3.92 0.1046
X1X3 43.21 1 43.21 8.27 0.0348
X2 X3 1.09 1 1.09 0.21 0.6673
X12 100.59 1 100.59 19.25 0.0071
X22 229.67 1 229.67 43.95 0.0012
X32 38.42 1 38.42 7.35 0.0422
Cor Total 1738.04 14 - - -

R2 = 0.9850, R2adj =0.9579, R2pre =0.7613

39
Table 5

Source Sum of DOF Mean F-value p-value > F


squares square
Model 1710.82 8 213.85 47.15 < 0.0001
X1-pH 852.15 1 852.15 187.87 < 0.0001
X2-Cn 57.15 1 57.15 12.60 0.0121
X3-C0 347.68 1 347.68 76.65 0.0001
X1X2 20.48 1 20.48 4.52 0.0777
X1X3 43.21 1 43.21 9.53 0.0215
X21 100.59 1 100.59 22.18 0.0033
X22 229.67 1 229.67 50.63 0.0004
X23 38.42 1 38.42 8.47 0.0270
Residual 27.22 6 4.54 - -
Lack of Fit 26.98 4 6.75 58.41 0.0169
Pure Error 0.23 2 0.12 - -
Cor Total 1738.04 14 - - -

R2 = 0.9843, R2adj = 0.9635, R2pre = 0.8225.

40

Vous aimerez peut-être aussi