You are on page 1of 8

Journal of Geochemical Exploration 147 (2014) 151–158

Contents lists available at ScienceDirect

Journal of Geochemical Exploration


journal homepage: www.elsevier.com/locate/jgeoexp

Optimization of heavy metal removal from aqueous solutions by


maghemite (γ-Fe2O3) nanoparticles using response
surface methodology
Ali Ahmadi a,b,⁎, Shahriar Heidarzadeh a, Ahmad Reza Mokhtari a,
Emaeel Darezereshki c, Houshang Asadi Harouni a
a
Department of Mining Engineering, Isfahan University of Technology, Isfahan, Iran
b
Mineral Bioprocessing Research Group (MBRG), Biotechnology and Bioengineering Research Institute, Isfahan University of Technology, Isfahan, Iran
c
Energy & Environmental Engineering Research Center, Shahid Bahonar University of Kerman, Kerman, Iran

a r t i c l e i n f o a b s t r a c t

Article history: The presence of heavy metals in the wastewaters is a serious threat to the environment. In this research, the re-
Received 6 February 2014 moval of some heavy metal ions, including Pb(II), Zn(II) and Cd(II) using maghemite (γ-Fe2O3) nanoparticles as
Received in revised form 13 October 2014 adsorbent was studied in an aqueous system. Experiments were carried out at a laboratory scale in 100 mL
Accepted 18 October 2014
Erlenmeyer flasks at 150 rpm for 120 min. A response surface methodology (RSM) with a Box–Behnken Design
Available online 27 October 2014
was employed to evaluate the effects of solution pH, temperature, initial heavy metal concentration and adsor-
Keywords:
bent dosage on the removal efficiency of the heavy metals. Results of analysis of variance (ANOVA) showed
Heavy metal removal that the initial metal concentration and pH were the most significant parameters for Cd(II) ion removal. Adsor-
Maghemite nanoparticles bent dosage, pH and their interaction effect had significant influences on the removal efficiency of Pb(II) ions. The
Optimization interaction effect of pH and initial concentration of metals had the most significant influences on the removal ef-
Adsorption ficiency of Zn(II) ions. The optimum pH, temperature, initial concentration of metals and adsorbent dosage were
found to be 7.5, 45 °C, 50 mg/L, and 4.74 g/L, respectively. The observed selectivity order was as follows: Pb(II)
(10.55 mg/g) N Zn(II) (4.79 mg/g) N (1.75 mg/g) Cd(II). It was confirmed from SEM-EDAX analyses that heavy
metal ions were present on the surface of nanoparticles after adsorption. Results indicated that maghemite nano-
particles can be used as an effective adsorbent for effluent decontamination especially in Pb–Zn bearing
wastewaters.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction Several conventional techniques have been employed to remove


heavy metal ions from wastewater effluents including but not limited
Heavy metals are toxic and unlike most organic contaminants, they to ion exchange, chemical precipitation, membrane processes, electro-
are not biodegradable in the environment (Feng et al., 2010; Fu and dialysis, reverse osmosis, electrochemical treatment and adsorption
Wang, 2011; Shen et al., 2009). They have a tendency to accumulate (Feng et al., 2010; Nassar, 2010; Shen et al., 2009; Song et al., 2011).
in living organisms and may cause serious harms and fatal health Most of these methods suffer from some disadvantages such as high
problems to plants, animals and human beings through food chain capital and operational costs for the treatment process and sludge
transfers (Feng et al., 2010; Fu and Wang, 2011; Shen et al., 2009; disposal (Nassar, 2010; Tuutijärvi et al., 2009). Among all of these ap-
Song et al., 2011). proaches, the adsorption process is considered more efficient and eco-
With the rapid development of industrial activities, especially in de- nomical because of its flexibility in design, simplicity of operation,
veloping countries, wastewaters contaminated with toxic heavy metals facile handling, and in many cases generation of high-quality treated ef-
are discharged directly or indirectly into the environment (Dhakal et al., fluent (Han et al., 2013; Hua et al., 2012). Therefore, many efforts have
2005; Fu and Wang, 2011; Shen et al., 2009). Cadmium (Cd) and lead been made for the development of low-cost and ample metal sorbents,
(Pb) have been considered as the most hazardous pollutant elements. such as activated carbon (Kobya et al., 2005), zeolite (Erdem et al., 2004;
Excess amounts of Zn in the body can also cause major health problems. Shawabkeh et al., 2004), chitosan (Juang and Shao, 2002), sawdust
These heavy metals are released from the Pb–Zn mine sites. (Shukla and Pai, 2005), functionalized silica (Bois et al., 2003), live bio-
mass (Yan and Viraraghavan, 2003) and agricultural byproducts
(Inbaraj and Sulochana, 2004; Nasiruddin Khan and Farooq Wahab,
⁎ Corresponding author at: Department of Mining Engineering, Isfahan University of
Technology, Isfahan PO Box. 84156-83111, Iran. Tel.: +98 311 391 5113; fax: +98 311
2007). These methods, however, have some drawbacks; including the
391 5176. generation of a large amount of waste sludge, the high regeneration
E-mail address: a.ahmadi@cc.iut.ac.ir (A. Ahmadi). cost, low adsorption rate and available capacity because of the existence

http://dx.doi.org/10.1016/j.gexplo.2014.10.005
0375-6742/© 2014 Elsevier B.V. All rights reserved.
152 A. Ahmadi et al. / Journal of Geochemical Exploration 147 (2014) 151–158

of intra-particle diffusion in highly porous adsorbents (Hu et al., 2006;


Shen et al., 2009). An effective adsorbent should generally possess a
high surface area, small diffusion resistance and short adsorption
equilibrium time, so that it can be used to remove larger amounts of
contaminants in a shorter time. In addition, it should generate a
minimum amount of sludge (Hu et al., 2006).
In recent years with the development of nanotechnology, numerous
types of magnetic nanoparticles have been successfully synthesized and
applied for the removal of heavy metals from aqueous solutions
(Afkhami and Moosavi, 2010; Hakami et al., 2012; Tang and Lo, 2013).
The magnetic nanoparticles possess the merits of large surface area
with a high number of surface active sites which result in increasing
both the rate and extent of heavy metal adsorption (Afkhami and
Moosavi, 2010; Hakami et al., 2012; Tang and Lo, 2013). The separation
of the adsorbent from solution would be so facile via magnetic field (Hu
et al., 2006). Maghemite nanoparticle as a novel adsorbent is expected
to offer an attractive and inexpensive option for the removal of heavy
metals (Hu et al., 2006). The main advantages of using maghemite
nanoparticles are: high adsorption capacity due to its large surface
area, simple and fast separation of metal-loaded adsorbent from
aqueous solutions using an external magnetic field, non-production of
contaminants, and easy synthesis by different methods (Hu et al.,
Fig. 1. TEM image of γ-Fe2O3 magnetic nanoparticles.
2006; Srinivasan and Viraraghavan, 2010).
Response surface methodology (RSM) is a set of mathematical and
statistical approaches used for designing, improving and optimizing
processes (Anderson and Whitcomb, 2005; Wantala et al., 2012). This 2.2. Adsorption experiments
method can be utilized for evaluating the effects of individual parame-
ters, their relative significance and the interaction of two or more The adsorption of heavy metals on maghemite nanoparticles was
variables and determining the optimum conditions for desired re- studied by a batch technique. A certain amount of adsorbent was equil-
sponses (Demim et al., 2014; Wantala et al., 2012). The main objective ibrated with 50 mL of the heavy metals solution of known concentration
of RSM is to determine the optimum operating conditions for the sys- in 100 mL Erlenmeyer flasks. The flasks were agitated in an incubator
tem or to determine a region that satisfies the operating specifications shaker at 150 rpm for 120 min. After that, the suspension of the adsor-
(Hammari et al., 2004; Mourabet et al., 2012; Myers et al., 2009). The bent was separated from the solution by a 0.22 μm cellulosic filter. The
RSM approach usually includes the following two steps: the first step concentration of heavy metal ions remaining in solution was measured
is the model formulation to determine which factors and their interac- by AAS (Perkin-Elmer 700) using the flame method. The effect of several
tions significantly affect the response, and the second step is the optimi- parameters, such as pH, temperature, contaminant concentration and
zation of the factors that influence the performance of the response adsorbent dose on the adsorption was investigated. The pH of the
(Stekelenburg et al., 2008). adsorptive solutions was adjusted using nitric acid (1 M) and sodium
In this study, the effects of some physicochemical parameters, i.e. hydroxide (1 M) when needed.
pH, temperature, initial heavy metal concentration and adsorbent
dosage on the adsorption efficiency of Pb(II), Cd(II) and Zn(II) ions 2.3. Design of experiments
on maghemite (γ-Fe2O3) nanoparticles were investigated. Using a
Box–Behnken Design, statistically designed experiments were used to Response surface methodology (RSM) was used to evaluate the ef-
determine variables that affect the heavy metal removal efficiency fects of different operating parameters on the removal efficiency of
more significantly. Attempts have also been made to optimize the re- Pb(II), Zn(II) and Cd(II) ions. It not only shows the optimum conditions,
moval process by using response surface optimization techniques. but it also proposes fitted regression models. A 3-level, 4-factor Box–
Behnken Design (BBD) was used to evaluate the effect of the selected

2. Materials and methods

2.1. Materials

Maghemite nanoparticles with a Fe3O4 content in excess of 99.5%


were prepared by a wet chemical method previously described
(Darezereshki et al., 2010). The average particle size of the maghemite
was estimated by transmission electron microscope (TEM) to be around
14 nm. Fig. 1 shows the TEM image of the particles and Fig. 2 shows
the FTIR wave numbers for maghemite to be 454.6, 587.55, and
632.83 cm−1.
Pb(II), Zn(II) and Cd(II) stock solutions were prepared by dissolv-
ing a specified weight of the related metal salts in 500 mL of deion-
ized water, and the working solutions with a desired concentration
were prepared by appropriate dilutions of the stock solution
immediately prior to their use. Pb(NO 3) 2 , Zn(NO 3) 2 ·4H2O and
Cd(NO 3) 2 ·4H2O were of analytical grade and supplied by Merck Fig. 2. FTIR-Spectra of γ-Fe2O3 magnetic nanoparticles performed by 200 (a), 250 (b) and
company (Germany). 300 (c) mg KBr.
A. Ahmadi et al. / Journal of Geochemical Exploration 147 (2014) 151–158 153

Table 1 The experimental results were analyzed using Design Expert 7.0 and
Actual and coded values of the test variables. a regression quadratic polynomial model was proposed as follows:
Factors Coded values −1 0 +1 X X X
2
pH 4.5 6 7.5
Yð%Þ ¼ α0 þ αi Xi þ αii Xi þ αi j Xi X j þ ε
Temperature (°C) 25 35 45
Contaminant concentration (mg/L) 10 30 50 where, α0 is the constant coefficient, αi, αii and αij are the regression co-
Adsorbent dose (g/L) 1 3 5
efficient and Xi, Xj indicate the independent variables. ε represents the
random error.

parameters on the removal efficiency of Pb(II), Cd(II) and Zn(II) ions 2.4. Analyses
from aqueous solutions by maghemite nanoparticles. The four parame-
ters affecting heavy metal removal, namely solution pH (X1), tempera- Atomic absorption spectrophotometry was used for the analysis
ture (X2), initial contaminant concentration (X3) and adsorbent dosage of heavy metals in solution. Scanning electron microscopy (SEM)
(X4) were selected as independent variables, and the removal efficien- equipped with an energy dispersive X-ray microanalyzer (EDAX)
cies (Y) of Pb(II), Zn(II) and Cd(II) were considered as the dependent (Cam-Scan MV 2300) was used to study the surface of solid residues
variables (responses). Variables were coded in accordance with the after the adsorption process. Prior to the SEM studies, the samples
following equation: were coated with a thin layer of gold.

X i −X 0 3. Results and discussion


xi ¼ ð1Þ
ΔX i
Statistical analysis by response surface methodology (RSM) was
used to determine a well-fitted regression model of adsorption process.
where xi is the coded value of an independent variable, Xi is the real The experimental data of each adsorbate (Pb(II), Zn(II) and Cd(II)) were
value of an independent variable, X0 is the real value of an independent fitted with linear, interactive, quadratic and cubic models to get the re-
variable at the center point and ΔXi is the step change value (Zheng and gression equations. The significance of suggested regression models for
Wang, 2010). The removal efficiency of each contaminant (Yi) was each adsorbate could be determined through the ANOVA. The final em-
calculated according to the following equation: pirical equations obtained from experimental data which demonstrate
the relationship between contaminant removal efficiency and indepen-
C0 −Ct dent variables are given in Eqs. (4), (5) and (6) for Cd(II), Pb(II) and
Y¼  100 ð2Þ Zn(II), respectively.
C0
2
YCdðIIÞ ¼ 11:82 þ 2:45  A þ 4:84  C−2:56  C ð4Þ
where, Y is the heavy metal removal efficiency (%), and C0 and Ct are the
initial and residual (after 120 min) concentrations of metal in solution
(mg/L). The experimental range and levels of independent variables
for metal ion removal were given in Table 1. A set of 29 experimental YPbðIIÞ ¼ 76:20 þ 0:59  A−1:07  B þ 12:28  D−7:37  AD
2 2
runs, including duplicates were designed using BBD (Table 2). þ 5:68  A þ 6:51  B ð5Þ

Table 2
Box–Behnken Design matrix for four variables and response values.

Experimental run pH (A) Temperature (B) Metal concentration in solution (C) Adsorbent dosage (D) Cd removal (%) Pb removal (%) Zn removal (%)

1 1.00 0.00 0.00 1.00 20.77 99.33 52.10


2 −1.00 0.00 −1.00 0.00 1.00 69.90 0.50
3 1.00 1.00 0.00 0.00 11.63 95.27 22.10
4 1.00 0.00 −1.00 0.00 2.50 88.40 34.40
5 −1.00 0.00 1.00 0.00 12.84 69.70 76.42
6 1.00 0.00 1.00 0.00 22.76 95.92 36.50
7 0.00 1.00 0.00 1.00 17.30 94.33 14.67
8 0.00 0.00 0.00 0.00 12.43 65.63 1.37
9 0.00 −1.00 −1.00 0.00 14.30 95.80 37.30
10 0.00 0.00 0.00 0.00 13.40 71.70 7.83
11 0.00 1.00 1.00 0.00 10.24 83.00 16.94
12 0.00 0.00 0.00 0.00 11.00 76.77 8.03
13 −1.00 −1.00 0.00 0.00 9.07 78.80 11.50
14 0.00 −1.00 0.00 1.00 8.97 97.30 18.53
15 0.00 0.00 0.00 0.00 9.07 74.27 6.033
16 0.00 −1.00 1.00 0.00 13.76 84.42 18.22
17 1.00 −1.00 0.00 0.00 13.60 98.87 44.03
18 0.00 0.00 −1.00 −1.00 5.00 91.30 2.00
19 0.00 −1.00 0.00 −1.00 12.73 57.83 7.87
20 0.00 0.00 0.00 0.00 10.00 74.43 6.60
21 0.00 0.00 −1.00 1.00 4.30 91.00 2.00
22 −1.00 1.00 0.00 0.00 13.50 92.63 14.30
23 1.00 0.00 0.00 −1.00 17.80 90.03 56.10
24 0.00 0.00 1.00 −1.00 11.76 73.14 17.40
25 0.00 1.00 −1.00 0.00 3.00 81.30 1.00
26 0.00 1.00 0.00 −1.00 11.53 56.67 1.90
27 0.00 0.00 1.00 1.00 14.94 89.58 18.28
28 −1.00 0.00 0.00 −1.00 11.13 48.47 2.43
29 −1.00 0.00 0.00 1.00 7.27 87.27 6.57
154 A. Ahmadi et al. / Journal of Geochemical Exploration 147 (2014) 151–158

Table 3 Table 5
ANOVA results of the regression model for optimization of Cd adsorption (determination ANOVA results of the regression model for optimization of Zn adsorption (determination
coefficient (R2) = 0.82; adjusted determination coefficient (R2adj) = 0.80). coefficient (R2) = 0.84; adjusted determination coefficient (R2adj) = 0.79).

Source Coefficient Sum of Degree of Mean F-value p-Value Source Coefficient Sum of Degree of Mean F-value p-Value
estimate squares freedom square (prob N F) estimate squares freedom square (prob N F)

Model 11.82 399.35 3 133.12 38.81 b0.0001 Model 0.86 7.04 6 1.17 18.14 b0.0001
A 2.45 71.80 1 71.8 20.93 0.0001 A 0.39 1.80 1 1.80 27.89 b0.0001
C 4.84 281.49 1 281.49 82.06 b0.0001 B −0.21 0.53 1 0.53 8.20 0.0090
C2 −2.56 46.06 1 46.06 13.43 0.0012 C 0.43 2.20 1 2.20 33.94 b0.0001
Residual 85.76 25 3.43 AC −0.54 1.16 1 1.16 18.00 0.0003
Lack-of-fit 77.51 21 3.69 1.79 0.3049 BC 0.36 0.51 1 0.51 7.96 0.0099
Pure error 8.25 4 2.06 A2 0.34 0.83 1 0.83 12.84 0.0017
Cor totala 485.11 28 Residual 1.42 22 0.065
a Lack-of-fit 1.35 18 0.075 3.95 0.0962
Total sum of squares corrected for the mean.
Pure error 0.076 4 0.019
Cor total 8.46 28

YZnðIIÞ ¼ 0:86 þ 0:39  A−0:21  B þ 0:43  C−0:54  AC þ 0:36


2
 BC þ 0:34  A ð6Þ
of the data is determined by its p-value being closer to zero (Srinivasan
where, YCd(II), YPb(II) and YZn(II) are the predicted responses for the re- and Viraraghavan, 2010). The main effect of each factor and the interac-
moval efficiency of Cd(II), Pb(II) and Zn(II) ions, respectively, and A, B, tion effects are statistically significant when p-value is less than 0.05
C and D are the coded values of pH, temperature, initial contaminant (Demim et al., 2014). As can be seen from Table 3, all the p-values of
concentration and adsorbent dosage, respectively. A, C and C2 are less than 0.05, which indicates that these variables are
Results of ANOVA for the removal of Cd(II), Pb(II) and Zn(II) were significant on the removal of Cd(II). Based on this hypothesis, A, D, AD,
tabulated in Tables 3, 4 and 5, respectively. The Fisher variation ratio A2 and B2 are the significant variables on the removal of Pb(II) and A,
(F-value) is a statistically valid measure of how well the factors describe B, C, AC, BC and A2 are the significant variables on the removal of Zn(II).
the variation in the data about its mean. The regression models were Response surface plots were used to determine the individual and
statistically significant with a confidence level of 95% at an calculated interaction effects of the input variables on the target responses. They
F-value of 38.8 for Cd(II), 19.5 for Pb(II) and 18.1 for Zn(II) with a very analyze the geometric nature of the surface, the maxima and minima
low probability value (p-value ≤ 0.0001) for all three heavy metal of the response and the significance of the coefficients of the equation
ions. The tabulated F-value for Cd(II) (F0.05, 3, 27), Pb(II) (F0.05, 3, 27) (Prakash et al., 2008).
and Zn(II) (F0.05, 3, 27) were found to be 2.29, 2.34 and 2.34, respectively.
Since, the calculated F-values for all the three models were higher than
3.1. Cadmium removal
the related tabulated F-values (38.8 N 2.29; 19.5 N 2.34; 18.1 N 2.34),
they have statistical significance. “Adequate Precision” ratios were
The response surface plots were obtained by varying two factors
21.19, 17.2 and 15.63 for Cd(II), Pb(II) and Zn(II) removal models re-
while keeping the other constant. Fig. 3 represents the effect of initial
spectively, which were far greater than the minimum desirable amount
concentration of metals and pH on Cd(II) removal. It shows that the ini-
of 4.0 which indicate the presence of adequate signal to noise for all
tial Cd(II) concentration had a positive effect on the removal efficiency
models.
of Cd(II) ions (the coefficient of C in Eq. (4) = 4.84), in which by in-
The value of determination coefficient (R2) determines the
creasing the initial concentration of Cd(II) from 10.0 to 50.0 mg/L, the
goodness-of-fit of the models. The adjusted R2 (R2adj) and predicted R2
Cd(II) removal efficiency increased from 4.4% to 14.1%; however, ac-
should be within approximately 0.20 of each other to be in reasonable
cording to Eq. (4), increasing the initial concentration of Cd(II) too
agreement. The close correspondence between R2adj and R2 indicates
much, had a negative effect on the removal efficiency (the coefficient
that unnecessary variables have not been included.
of C2 in Eq. (4) = − 2.56). This behavior could be related to the
As shown in Tables 3 to 5, the R2 values of 0.82, 0.85 and 0.84 respec-
tively, for Cd(II), Pb(II) and Zn(II) removal models showed the well
fitness of regression models for predicting the removal results.
The regression coefficient values of independent variables of
Eqs. (4) to (6) were listed in Tables 3 to 5. p-Value is used to determine
the effects in the model that are statistically significant. The significance

Table 4
ANOVA results of the regression model for optimization of Pb adsorption (determination
coefficient (R2) = 0.85; adjusted determination coefficient (R2adj) = 0.8).

Source Coefficient Sum of Degree of Mean F-value p-Value


estimate squares freedom square (prob N F)

Model 76.2 3839.89 6 639.98 19.5 b0.0001


A 10.59 1345.21 1 1345.21 40.99 b0.0001
B −1.07 13.72 1 13.72 0.42 0.5246
D 12.28 1809.91 1 1809.91 55.16 b0.0001
AD −7.37 217.56 1 217.56 6.63 0.0173
A2 5.68 222.56 1 222.56 6.78 0.0162
B2 6.51 292.42 1 292.42 8.91 0.0068
Residual 594.12 22 32.81
Lack-of-fit 521.29 18 36.06 1.98 0.2674
Pure error 72.84 4 18.21
Fig. 3. 3D plots for interactive effect of pH and initial Cd(II) concentration on the removal
Cor total 4561.81 28
efficiency of Cd(II) at temperature = 35 °C and adsorbent dosage = 3 g/L.
A. Ahmadi et al. / Journal of Geochemical Exploration 147 (2014) 151–158 155

complexity of competitive adsorption of the metal ions at different con-


centrations on the surface of the nanoparticles. In the higher range of
adsorbate concentration, the surface of adsorbent began to get saturated
and results in decreasing the adsorption efficiency of Cd(II) ions. As
shown in Fig. 3, solution pH was the other significant factor and showed
a positive effect (the coefficient of A in Eq. (4) = 2.45). The removal ef-
ficiency of Cd(II) increased from 9.37% to 14.26% by increasing the pH
from 4.5 to 7.5. The pH of the zero point of charge (pHzpc) value for
maghemite is about 6.1. Electrostatic attraction is considered as the
main adsorption mechanism at pH values higher than the pHzpc. Metal
ions can also adsorb on the surface of iron oxides through the ion ex-
change process at pH values less than the pHzpc (the exchange of proton
with cation) (Tang and Lo, 2013).
The low removal efficiency of Cd(II) at pH values below the pHzpc is
attributed to the positively charged of the maghemite surface. The re-
moval of Cd(II) at the pH values lower than the pHzpc, is justified by
the exchange of adsorbed proton with Cd(II) ions.

3.2. Lead removal Fig. 5. 3D plots for interactive effect of pH and adsorbent dosage at initial Pb(II) concentra-
tion = 30 mg/L and temperature = 35 °C on the adsorption efficiency of Pb(II).
Figs. 4 and 5 show the effects of significant variables on the removal
efficiency of Pb(II) ions. The removal efficiency of Pb(II) on maghemite
nanoparticles was very sensitive to the changes in solution pH (Coeffi-
dosage parameters individually had a positive effect on the removal ef-
cient = 10.59 in Eq. (6)). It shows that increasing in pH value from
ficiency of Pb(II) ions, the simultaneous increase of these parameters
4.5 to 7.5 resulted in enhancing the removal efficiency of Pb(II) ions
showed a negative synergistic effect on the removal.
from 71.29% to 92.47% (Fig. 4). It can also be pointed out that at the
pH values lower than the pHzpc, some Pb(II) ions can be adsorbed on
the surface of maghemite nanoparticles through the ion exchange pro- 3.3. Zinc removal
cess. It was reported (Lodeiro et al., 2006; Naiya et al., 2008) that the
precipitation of lead hydroxide can play a leading role during heavy Fig. 6 shows a response surface plot where Zn(II) removal was
metal removal process at pH above 6. The effect of temperature and represented by varying simultaneously pH from 4.5 to 7.5 and Zn con-
pH on the removal efficiency of Pb(II) ions is represented in Fig. 4. As centration from 10 to 50 mg/L. From this plot, it can be found that
can be seen, increasing temperature too much, had a significant positive about 58% Zn removal can be achieved at the pH value of 4.5 and the
effect on the removal efficiency of Pb(II) ions (coefficient = 6.51 in Zn concentration of 50 mg/L. The removal was sensitive to the interac-
Eq. (6)). The increased removal of Pb(II) ions with increasing tempera- tion effect of pH and the initial concentration of Zn(II) which had a neg-
ture may be a result of the faster chemical precipitation rate of lead hy- ative synergistic effect on the target variable (Figs. 6 and 7). However,
droxide at the higher temperatures. The interactive effects of pH and increasing pH from 4.5 to 7.5 and initial Zn(II) concentration from
adsorbent dosage were shown in Fig. 5. They show that increasing in ad- 10.0 to 50.0 mg/L increased the removal efficiency of Zn(II) ions from
sorbent dosage has the highest effect on the adsorption of Pb(II) ions 6.6% to 39.1% and from 2.7% to 19.1%, respectively. At the higher level
(coefficient = 12.28 in Eq. (6)). By increasing the adsorbent dosage of pH, the removal efficiency of Zn(II) was increased mainly as a result
from 1.0 (g/L) to 5.0 (g/L), the removal efficiency of Pb(II) ions rose of electrostatic attraction. The removal of Zn(II) at pH values less than
from 63.92% to 88.48%. This behavior could be ascribed to a greater sur- the pHzpc is related to the cation exchange mechanism. This assumption
face area and the more availability of adsorption sites at the higher ad- was supported by the decreasing of the solution pH during the process
sorbent dosage. Although the increase in each of pH and adsorbent which could be a result of releasing protons into the solution. Fig. 7

Fig. 4. 3D plots for interactive effect of pH and temperature at initial Pb(II) concentration = Fig. 6. 3D plots for interactive effect of pH and initial Zn concentration at temperature =
30 mg/L and adsorbent dosage = 3 g/L. 35 °C and adsorbent dosage = 3 g/L on the removal efficiency of Zn(II).
156 A. Ahmadi et al. / Journal of Geochemical Exploration 147 (2014) 151–158

found to be at the initial solution pH of 7.5, the temperature of 45 °C,


the initial ion concentration of 50 mg/L and the adsorbent (dosage of
4.74 g/L). In this condition the removal efficiencies of Cd(II), Pb(II) and
Zn(II) ions were 16.6%, 100.0% and 42.5%, respectively and the desirabil-
ity value was found to be 0.90. This optimum condition was checked ex-
perimentally. The results showed the removal efficiency of 16.2%,
100.0% and 42.1% for Cd(II), Pb(II) and Zn(II) ions, respectively. The
high degree of agreement between the predicted optimum conditions
and the repeated experimental results indicated that the Box–Behnken
Design could be employed as an effective and reliable tool for evaluation
and optimization of the effects of adsorption parameters on the removal
efficiency of heavy metals using maghemite nanoparticles.
SEM-EDAX analysis was used to characterize the chemical proper-
ties and morphology of the adsorbent. The results were presented in
Fig. 9. It was confirmed from this analysis that the heavy metal ions
were present on the surface of nanoparticles after the adsorption
process.

Fig. 7. 3D plots for interactive effect of temperature and initial Zn concentration in solution
at pH = 6 and adsorbent dosage = 3 g/L on the adsorption efficiency of Zn(II). 4. Conclusion

showed the positive effects of temperature and initial Zn(II) concentra- In this research, maghemite nanoparticles were used to remove
tion on the heavy metal removal efficiency. It shows that the initial con- some heavy metal ions, i.e. Pb(II), Zn(II) and Cd(II), from aqueous solu-
centration of the metal is more significant than temperature under the tions. Response surface methodology was used to find a maximum loca-
high concentration of Zn(II) ions. tion in the design space. A Box–Behnken Design was employed to
evaluate the effects of pH, temperature, initial contaminant concentra-
tion and adsorbent dosage on the removal efficiency of Pb(II), Zn(II)
3.4. Optimization and Cd(II) ions on maghemite (γ-Fe2O3) nanoparticles under competi-
tive conditions. Quadratic models were well fitted to the experimental
Maximum removal efficiency for each adsorbate and the corre- data and second-order polynomial equations (regression models)
sponding optimal conditions of variables were determined and the were described the relationship between the responses and the vari-
models were confirmed by some further experimental runs. Numerical ables accurately. Results showed that the initial concentration of
optimization was done to find a maximum point for the desirability Cd(II) and pH were the most significant parameters on the removal
function (Mourabet et al., 2012) by setting the values of pH, tempera- efficiency of Cd(II) ions. Adsorbent dosage, pH and their interaction ef-
ture, concentration of initial ions and adsorbent dosage within their fects had the most significant influences on the removal efficiency of
ranges and maximizing the removal efficiencies of Cd(II), Pb(II) and Pb(II) ions. Moreover, the interaction effect of pH and the initial concen-
Zn(II) ions. As can be seen in Fig. 8, the best local maximum value was tration Pb(II) had the most significant influences on the removal

Fig. 8. Desirability ramp for numerical optimization of four independent variables and the responses.
A. Ahmadi et al. / Journal of Geochemical Exploration 147 (2014) 151–158 157

Fig. 9. SEM/EDAX photomicrographs of maghemite nanoparticles after adsorption process at the optimum conditions.

efficiency of Zn(II). The predicted maximum adsorption capacities were Demim, S., Drouiche, N., Aouabed, A., Benayad, T., Couderchet, M., Semsari, S., 2014. Study
of heavy metal removal from heavy metal mixture using the CCD method. J. Ind. Eng.
10.55, 4.49 and 1.75 mg/g for Pb(II), Zn(II) and Cd(II), respectively. The Chem. 20, 512–520.
corresponding optimal conditions of variables of adsorption process Dhakal, R.P., Ghimire, K.N., Inoue, K., 2005. Adsorptive separation of heavy metals from an
were determined to be at the pH of 7.5, the temperature of 45 °C, the aquatic environment using orange waste. Hydrometallurgy 79, 182–190.
Erdem, E., Karapinar, N., Donat, R., 2004. The removal of heavy metal cations by natural
initial metal concentration of 50 mg/L (each metal ion) and the adsor- zeolites. J. Colloid Interface Sci. 280, 309–314.
bent dosage of 4.74 g/L. The removal efficiency of heavy metal ions by Feng, Y., Gong, J.-L., Zeng, G.-M., Niu, Q.-Y., Zhang, H.-Y., Niu, C.-G., Deng, J.-H., Yan, M.,
the nanoparticles was found to follow the order of Pb N Zn N Cd. 2010. Adsorption of Cd (II) and Zn (II) from aqueous solutions using magnetic
hydroxyapatite nanoparticles as adsorbents. Chem. Eng. J. 162, 487–494.
SEM-EDAX analysis confirmed that the heavy metal ions were
Fu, F., Wang, Q., 2011. Removal of heavy metal ions from wastewaters: a review. J. Environ.
present on the surface of nanoparticles after adsorption process. Manag. 92, 407–418.
The results of this study suggested that maghemite nanoparticles Hakami, O., Zhang, Y., Banks, C.J., 2012. Thiol-functionalised mesoporous silica-coated
magnetite nanoparticles for high efficiency removal and recovery of Hg from water.
could be used as a potential and appealing adsorbent for the removal
Water Res. 46, 3913–3922.
of heavy metals from aqueous solutions. Hammari, L., Laghzizil, A., Barboux, P., Lahlil, K., Saoiabi, A., 2004. Retention of fluoride
ions from aqueous solution using porous hydroxyapatite: structure and conduction
properties. J. Hazard. Mater. 114, 41–44.
Han, C., Pu, H., Li, H., Deng, L., Huang, S., He, S., Luo, Y., 2013. The optimization of
References As (V) removal over mesoporous alumina by using response surface methodology
and adsorption mechanism. J. Hazard. Mater. 254, 301–309.
Afkhami, A., Moosavi, R., 2010. Adsorptive removal of Congo red, a carcinogenic textile Hu, J., Chen, G., Lo, I.M., 2006. Selective removal of heavy metals from industrial wastewater
dye, from aqueous solutions by maghemite nanoparticles. J. Hazard. Mater. 174, using maghemite nanoparticle: performance and mechanisms. J. Environ. Eng. 132,
398–403. 709–715.
Anderson, M.J., Whitcomb, P.J., 2005. RSM Simplified: Optimizing Processes Using Re- Hua, M., Zhang, S., Pan, B., Zhang, W., Lv, L., Zhang, Q., 2012. Heavy metal removal from
sponse Surface Methods for Design of Experiments. Productivity Press. water/wastewater by nanosized metal oxides: a review. J. Hazard. Mater. 211, 317–331.
Bois, L., Bonhommé, A., Ribes, A., Pais, B., Raffin, G., Tessier, F., 2003. Functionalized silica for Inbaraj, B.S., Sulochana, N., 2004. Carbonised jackfruit peel as an adsorbent for the removal
heavy metal ions adsorption. Colloids Surf. A Physicochem. Eng. Asp. 221, 221–230. of Cd (II) from aqueous solution. Bioresour. Technol. 94, 49–52.
Darezereshki, E., Ranjbar, M., Bakhtiari, F., 2010. One-step synthesis of maghemite Juang, R.-S., Shao, H.-J., 2002. A simplified equilibrium model for sorption of heavy metal
(γ-Fe2O3) nano-particles by wet chemical method. J. Alloys Compd. 502, 257–260. ions from aqueous solutions on chitosan. Water Res. 36, 2999–3008.
158 A. Ahmadi et al. / Journal of Geochemical Exploration 147 (2014) 151–158

Kobya, M., Demirbas, E., Senturk, E., Ince, M., 2005. Adsorption of heavy metal ions from Shukla, S., Pai, R.S., 2005. Removal of Pb (II) from solution using cellulose‐containing
aqueous solutions by activated carbon prepared from apricot stone. Bioresour. materials. J. Chem. Technol. Biotechnol. 80, 176–183.
Technol. 96, 1518–1521. Song, J., Kong, H., Jang, J., 2011. Adsorption of heavy metal ions from aqueous solution by
Lodeiro, P., Herrero, R., Vicente, M., 2006. The use of protonated Sargassum muticum as polyrhodanine-encapsulated magnetic nanoparticles. J. Colloid Interface Sci. 359,
biosorbent for cadmium removal in a fixed-bed column. J. Hazard. Mater. 137, 244–253. 505–511.
Mourabet, M., El Rhilassi, A., El Boujaady, H., Bennani-Ziatni, M., El Hamri, R., Taitai, A., Srinivasan, A., Viraraghavan, T., 2010. Oil removal from water by fungal biomass: a
2012. Removal of fluoride from aqueous solution by adsorption on hydroxyapatite factorial design analysis. J. Hazard. Mater. 175, 695–702.
(HAp) using response surface methodology. J. Saudi Chem. Soc. (in press). Stekelenburg, D.T., Lukszo, Z., Lowe, J., 2008. Improvement of the production process of
Myers, R.H., Montgomery, D.C., Anderson-Cook, C.M., 2009. Response Surface Methodology: leached optical fibers in a technological and organizational context. Comput. Aided
Process and Product Optimization Using Designed Experiments. John Wiley & Sons. Chem. Eng. 25, 1027–1032.
Naiya, T., Bhattacharya, A., Das, S., 2008. Adsorption of Pb (II) by sawdust and neem bark Tang, S.C., Lo, I., 2013. Magnetic nanoparticles: essential factors for sustainable environ-
from aqueous solutions. Environ. Prog. 27, 313–328. mental applications. Water Res. 47, 2613–2632.
Nasiruddin Khan, M., Farooq Wahab, M., 2007. Characterization of chemically modified Tuutijärvi, T., Lu, J., Sillanpää, M., Chen, G., 2009. As (V) adsorption on maghemite nano-
corncobs and its application in the removal of metal ions from aqueous solution. J. particles. J. Hazard. Mater. 166, 1415–1420.
Hazard. Mater. 141, 237–244. Wantala, K., Khongkasem, E., Khlongkarnpanich, N., Sthiannopkao, S., Kim, K.-W., 2012.
Nassar, N.N., 2010. Rapid removal and recovery of Pb (II) from wastewater by magnetic Optimization of As (V) adsorption on Fe-RH-MCM-41-immobilized GAC using Box–
nanoadsorbents. J. Hazard. Mater. 184, 538–546. Behnken Design: effects of pH, loadings, and initial concentrations. Appl. Geochem.
Prakash, O., Talat, M., Hasan, S., Pandey, R.K., 2008. Factorial design for the optimization of 27, 1027–1034.
enzymatic detection of cadmium in aqueous solution using immobilized urease from Yan, G., Viraraghavan, T., 2003. Heavy-metal removal from aqueous solution by fungus
vegetable waste. Bioresour. Technol. 99, 7565–7572. Mucor rouxii. Water Res. 37, 4486–4496.
Shawabkeh, R., Al-Harahsheh, A., Hami, M., Khlaifat, A., 2004. Conversion of oil shale ash Zheng, Y., Wang, A., 2010. Removal of heavy metals using polyvinyl alcohol semi-IPN poly
into zeolite for cadmium and lead removal from wastewater. Fuel 83, 981–985. (acrylic acid)/tourmaline composite optimized with response surface methodology.
Shen, Y., Tang, J., Nie, Z., Wang, Y., Ren, Y., Zuo, L., 2009. Preparation and application of Chem. Eng. J. 162, 186–193.
magnetic Fe3O4 nanoparticles for wastewater purification. Sep. Purif. Technol. 68,
312–319.

You might also like