Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Indian Chemical Engineer

ISSN: 0019-4506 (Print) 0975-007X (Online) Journal homepage: http://www.tandfonline.com/loi/tice20

Statistical Optimization of Electrocoagulation


Process for Removal of Nitrates Using Response
Surface Methodology

Sanigdha Acharya, S.K. Sharma, Garima Chauhan & Darshan Shree

To cite this article: Sanigdha Acharya, S.K. Sharma, Garima Chauhan & Darshan Shree (2017):
Statistical Optimization of Electrocoagulation Process for Removal of Nitrates Using Response
Surface Methodology, Indian Chemical Engineer, DOI: 10.1080/00194506.2017.1365630

To link to this article: http://dx.doi.org/10.1080/00194506.2017.1365630

Published online: 05 Sep 2017.

Submit your article to this journal

Article views: 5

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tice20

Download by: [GGS Indraprastha University] Date: 07 September 2017, At: 22:29
INDIAN CHEMICAL ENGINEER © 2017 Indian Institute of Chemical Engineers
https://doi.org/10.1080/00194506.2017.1365630

Statistical Optimization of Electrocoagulation


Process for Removal of Nitrates Using Response
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

Surface Methodology

Sanigdha Acharya*, S.K. Sharma, Garima Chauhan and Darshan Shree


University School of Chemical Technology, Guru Gobind Singh Indraprastha University, Dwarka, India

Abstract: The present study explores the applicability of response surface methodology in conjunction
with the Box–Behnken design (BBD) matrix to statistically optimize the electrocoagulation process for
the removal of nitrates from aqueous solution. To the best of our knowledge, the BBD matrix has not
been investigated yet in the literature for the removal of nitrate using the electrocoagulation process.
The synergetic effect of three process parameters, i.e. pH (7–12), electrolysis time (20–100 min) and
current (0.5–1.5 A), was studied. The quadratic regression model demonstrated a good correlation
between predicted and experimental values (R 2 = 0.95). Confirmatory experiments showed maximum ±
4% deviation in theoretical and experimental data. ANOVA studies (p-value< .05) confirmed the
adequacy of the regression model. In addition, a low value (5.56) of coefficient of variation indicated a
higher degree of precision and reliability of the experiments. Response surfaces demonstrated a
significant effect of electrolysis time and current on the removal of nitrate from aqueous solution,
whereas reaction pH does not show a remarkable change in removal efficiency. The optimum
conditions obtained were electrolysis time 77 min, current 1.2 A and pH 9.8 to achieve maximum
89.4% removal of nitrate. The proposed design correlations may prove to be a useful tool in designing
the pilot plant/commercial plant for the electrocoagulation process.

Keywords: Nitrates, Electrocoagulation, Aluminium electrodes, RSM, ANOVA, Box–Behnken design.

1. Introduction
Revolutionary industrialization and technological advancements have certainly made a
remarkable transformation in human life, but negative shades of industrial activities cannot
be denied. Leaching of chemical fertilizers from cultivated land into soil, heedless discharge

*Author for Correspondence. Email: sanigdha@rediffmail.com


2 ACHARYA ET AL .

of waste into the ecosystem and other anthropological activities increase the nitrate content in
the water bodies. Indeed, about 30–70% of nitrogen fertilizer used in agriculture is lost to the
environment and found in the form of nitrate in surface and groundwater [1,2]. The presence
of even small amount of nitrates (<10 mg/L) in ground and surface water may cause methae-
moglobinemia which can be fatal to neonates [3]. Nitrates, being a precursor of nitrosamines,
may be responsible for hypertension, thyroid disability, carcinogenic and mutagenic health pro-
blems [4]. Environmental regulations for nitrate concentration in drinking water are becoming
increasingly stringent worldwide. The acceptable limit of nitrates in drinking water is 50 mg/L
as prescribed by the World Health Organization [5], whereas it is 45 mg/L in India [6].
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

Nitrate, being a stable and highly soluble ion, is very difficult to remove using conventional
methods such as chemical coagulation, lime softening and oxidation processes [3,7]. A survey
of the literature yielded information on the advanced physicochemical and biological treatment
methods (biological denitrification [8–12], ion exchange [13–17], reverse osmosis [18–20] and
electrodialysis [21]) to remove nitrate from water. These methods have several limitations
such as long reaction time, high temperature sensitivity for biological treatment, limited selec-
tivity in the presence of competing anions and generation of concentrated brine at large scale
for ion-exchange methods, membrane fouling and operating cost for reverse osmosis and elec-
trodialysis [22]. Therefore, it is necessary to develop an effective treatment method for the
removal of nitrates from water. Several studies have been aimed at development of suitable cat-
alysts for the catalytic reduction of nitrate [23,24]. Various emerging adsorbents such as gra-
phene [25], activated carbon [26] and clay [27] have also been investigated for the removal of
nitrates from water. Recently, researchers have revealed electrocoagulation (EC) as an attractive
and suitable treatment method for various types of water and wastewaters because of its flexi-
bility, safety, ease of operation and cost effectiveness [28]. In EC, coagulant generation is done
in situ by applying current, which helps in the removal of the pollutants by floc formation and
sedimentation. There is formation of gas bubbles also, which entrain the lighter pollutants and
bring them to the surface of the solution and which are removed by flotation [29,30].
As demonstrated by most of the recent literature, EC is having a significant role for the
removal of organic pollutants [31], turbidity [32], dyes [33,34], pharmaceuticals [35] and
heavy metals [36]. Kamaraj and Vasudevan [37] investigated the efficiency of facile one-pot
electro-synthesized Al(OH)3 for the removal of organic pollutants from water and demon-
strated maximum 86% removal of 2,4,5-trichlorophenoxyacetic acid at a current density of
0.10 A/dm2 and a pH value of 7 [37]. Similarly, herbicide 2-(2,4-dichlorophenoxy) propanoic
acid was also removed from aqueous solution by anodic dissolution of sacrificial aluminium
anode [38]. Heffron et al. [39] investigated the effect of electrode material, post-treatment,
water composition and pH on the removal of trace metal (arsenic, cadmium, chromium,
lead and nickel) from drinking water by EC. Cheballah et al. [40] used bipolar EC for the sim-
ultaneous removal of hexavalent chromium and COD from industrial wastewater. The results
showed that this process could effectively remove 100% chromium (VI) and 95.95% of COD.
Similarly, bipolar EC was used for boron removal and 96% efficiency was reported [41].
Kamaraj and Vasudevan [42] investigated the efficiency of the EC process for the removal of
radioactive contaminants (thorium (Th4+), uranium (U4+) and cerium (Ce4+)) from aqueous
solution with the aid of zinc sacrificial anodes. Palahouane et al. [43] studied efficiency of
EC for fluoride removal from photovoltaic wastewater and reported optimum fluoride

INDIAN CHEMICAL ENGINEER 2017


Statistical Optimization of Electrocoagulation Process 3

removal (95%) at operating conditions: current density, 18.51 A/m2; number of electrodes of 3,
and initial pH 7. EC has also been successfully applied to remove inorganic species from
aqueous solutions [44,45] though few studies have been reported for the removal of nitrates
using EC. Vasudevan et al. [46] used EC for the removal of NO− 3 from drinking water using
magnesium electrodes and obtained 95.8% removal efficiency at a current density of 0.25 A/
dm2 and pH 7. Pak [45] compared performance of aluminium and iron electrodes for the
removal of nitrates using EC and found that aluminium electrodes were more efficient than
iron electrodes. Kamaraj et al. [47] used one-pot electrochemically synthesized nano-zinc
hydroxide for the adsorption of nitrate from aqueous solution. Effects of co-existing anions,
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

such as carbonate, phosphate, silicate and phosphate, were investigated and maximum
removal efficiency (69%) was obtained at neutral pH and a current density of 0.1 A/dm2.
Kumar and Goel [48] worked with the continuous flow reactor for the removal of nitrates
using the EC process and reported 84% removal efficiency at 25 V. Lakshmi et al. [49] used
the EC process for the simultaneous generation of hydrogen and removal of nitrate (NO− 3)
from water using aluminium alloy as the anode and cathode. It was concluded from the
study that energy yield of generated hydrogen was ∼54% of the electrical energy demand of
the EC process.
These studies follow the principle of “one variable at a time (OVAT)” approach in which
only one process parameter varies for a wide range, keeping all other parameters constant.
Therefore, the mutual interaction effect of process parameters on nitrate removal efficiency
cannot be estimated [50]. To overcome the limitations associated with OVAT, response
surface methodology (RSM) is receiving attention these days. RSM is a set of designed exper-
iments to obtain mathematical and statistical optimization for the multivariate analysis of
response. It is one of the suitable methods to optimize the operating conditions and maximizes
the pollutant removal [51]. Literature suggests that RSM has been successfully employed for
modelling and optimization of the organophosphorus compound degradation by photolysis
of hydrogen peroxide [52] and heavy metal removal using L. gibba [53,54]. Recently, it has
also been employed for optimization of process parameters for the removal of nitrate using
the EC process [55,56]. Bhanu and Prasad [57] conducted EC experiments using the central
composite design (CCD) matrix and optimized the interaction effect of current density, pH,
reaction time and distance between electrodes. CCD was also employed to predict the ecophy-
siological effects of a mixture of heavy metals on a Lemna gibba [58]. Another widely accepted
design matrix for coupling with RSM is the Box–Behnken design (BBD). BBD takes into
account axial and factorial points to get statistically designed combinations of independent
variable and thus, requires lesser number of experiments than CCD.
A series of experiments were carried out to remove nitrates from a synthetic solution by EC
and determine statistically the relationship among various process parameters using RSM. In
most electrochemical processes, current intensity, pH and electrolysis time are the most impor-
tant parameters for controlling the reaction rate [55,59]. Therefore, for the scope of this paper,
mutual interaction effect of these process variables on nitrate removal was investigated using
the BBD matrix. Other parameters, i.e. initial concentration, stirring speed and electrode dis-
tance, were kept constant. A mathematical regression model correlating the target response to
independent variables was developed and ANOVA studies were performed in order to

INDIAN CHEMICAL ENGINEER 2017


4 ACHARYA ET AL .

statistically analyse the multivariate response. 2D and 3D contours were plotted to identify the
optimized reaction conditions for efficient removal of nitrate from aqueous solutions.
The BBD matrix is an excellent technique to study the influence of major parameters on
response by considerably reducing the number of experiments and consequently saving time,
energy and cost. To the best of our knowledge, BBD has not been explored much for the
removal of nitrate using the EC process. Therefore, the major objective of the present study
is to investigate the applicability of the BBD matrix coupled with RSM for the EC process.
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

2. Material and Methods


2.1. Materials
Different industrial wastewaters (fertilizer, electroplating and sewage treatment) are reported to
contain more than 200 mg NO− 3 [60–62]. Therefore, a synthetic nitrate sample of 300 mg/L was
prepared by dissolving the appropriate amount of potassium nitrate (Qualikems, India) in dis-
tilled water. Conductivity of the solution was increased by adding NaCl (Thomas Baker, India).
The pH of the solution was adjusted by adding sodium hydroxide (Qualikems, India) or hydro-
chloric acid (Fisher Scientific, India).

2.2. Batch EC
Experiments were conducted in a batch mode of operation by taking 200 mL of the solution in
a glass beaker of capacity 250 mL. Aluminium electrodes were chosen in the present study for
the EC process. It is suggested in the literature that in situ generated Al(OH)3 showed an excel-
lent adsorption capacity and nitrate removal efficiency from aqueous solutions [37]. Alu-
minium electrodes of size 12.4 cm × 2.5 cm × 0.3 cm were used and spacing between the
electrodes was maintained at 1 cm. The electrodes were dipped up to 5 cm into the solution
and the effective area of the electrodes in the solution was 28.75 cm2. The schematic
diagram for the EC process is shown in Fig. 1.

Fig. 1. Schematic representation of EC process set-up.

INDIAN CHEMICAL ENGINEER 2017


Statistical Optimization of Electrocoagulation Process 5

To maintain uniform concentration, the contents of the beaker were agitated by a magnetic
stirrer (Labman, India). The electrodes were connected to a DC power supply (Kitheley, China)
providing 0–30 V and 0–3 A current. During the electrolysis process, aluminium hydroxides (Al
(OH)3) were formed rapidly by anodic dissolution of sacrificial aluminium electrodes as shown
in the following reactions (Equations (1)–(3)) [38]:

At the anode: Al  Al3+ + 3e− , (1)


Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

Al3+ (aq.) + 3H2 O  Al(OH)3 + 3H+ , (2)

At the cathode: 2H2 O + 2e−  H2 (g) + 2OH− . (3)

After the completion of the process, a retention time of 2 h was given and the sludge was
separated. Various parameters of the EC process are listed in Table 1 out of which three process
parameters, i.e. pH (7–12), electrolysis time (20–100 min) and current (0.5–1.5 A), were varied
for wide range. An amphoteric behaviour of Al(OH)3 leads to soluble Al3+ cations (at acidic
pH) which are not useful for water treatment. When the initial pH is kept neutral, polymeric
species are produced at the anode and precipitate as Al(OH)3 leading to more removal effi-
ciency [51]. Therefore, the effect of pH was investigated in the range of 7–12 in the present
study. Drouiche et al. [63] successfully used EC in treating solutions containing fluoride and
suggested that the quality of EC treated solution is closely dependent on applied potential
and electrolysis time. Therefore, current and electrolysis time were considered as variable
process parameters.
The concentration of nitrates in supernatant liquid was analysed using a UV–Vis spectro-
photometer (UV5704SS, India) at a wavelength of 220 nm. A linear relationship was observed
between the absorbance and concentration of nitrates in the solution. Nitrate removal

Table 1. Values of various parameters used for EC

Parameter Value
Fixed parameters
Initial NO−3 concentration (mg/L) 300
Stirring speed (rpm) 300
Electrode type Aluminium
Electrode distance (cm) 1
Effective area (cm2) 28.5
Retention time (min) 120
Variable parameters
pH 7–12
Electrolysis time (min) 20–100
Current (A) 0.5–1.5

INDIAN CHEMICAL ENGINEER 2017


6 ACHARYA ET AL .

efficiency was calculated using the following equation:

C0 − Ci
% removal = × 100. (4)
C0

where Co = initial nitrate concentration in mg/L and Ci = final nitrate concentration at any time
“t” in mg/L.

2.3. Design of Experiments


Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

Optimization of process parameters using RSM includes the investigation of statistically


designed combinations for maximizing the target response, estimation of coefficients by
fitting the experimental data to the response functions, prediction of response using the devel-
oped regression model and checking the adequacy of the model [64]. An independent BBD
matrix was used in the present study to design the batch electro-coagulation experiments.
The BBD matrix is a rotatable second-order design in which each independent variable is
placed at one of three equally spaced values [50]. BBD does not account for any embedded fac-
torial or fractional factorial points. It also allows calculations of the response functions at inter-
mediate levels and enables estimation of the system performance at any point within the
specified range of process parameters [64]. In the present study, three independent variables
(solution pH, current and electrolysis time) were selected to optimize the nitrate removal effi-
ciency. The total number of experiments for the BBD matrix was calculated using the following
equation:

N = 2k(k − 1) + Cp , (5)

where N = number of experiments; k = number of process variables; and Cp = centre points.


Totally, 17 experimental runs (4 axial, 8 factorial and 5-fold repetition of centre points) were
designed based on the BBD matrix at three different levels (−1, 0, 1) of independent parameters.
The target response (dependent variable) is nitrate removal efficiency (% Y ). Table 2 lists the
coded and experimental values of the independent variables.
RSM was employed in combination with the BBD matrix to evaluate the mutual inter-
action effect of independent variables on nitrate removal efficiency. Second-order multiple
regression analysis was employed to explain the behaviour of the system using the least-
squares regression methodology for obtaining the parameter estimators of the mathematical

Table 2. Coded and experimental values of process parameters.

Design matrix Parameter values pH Electrolysis time (min) Current (A)


(A) (B) (C)

BBD matrix Low level (−1) 7 20 0.5


Centre point (0) 9.5 60 1.0
High level (+1) 12 100 1.5

INDIAN CHEMICAL ENGINEER 2017


Statistical Optimization of Electrocoagulation Process 7

model [53]. The correlation between the nitrate removal efficiency (Y ) and the set of indepen-
dent variables (Xi , i = 1–3) was incurred by quadratic polynomial equation, as shown in the
following equation:


i=3 
i=3 
i=3 
i=3
Y = b0 + b i Xi + bii Xi2 + bij Xi Xj + e, (6)
i=1 i=1 i=1 i=j=1

where β0 = Constant; βi = Linear coefficient; βii = Squared coefficient; βij = Cross-product


Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

coefficient.
Analysis of variance (ANOVA) and calculation of regression coefficient were performed in
order to substantiate the adequacy of the proposed polynomial model. Three-dimensional
response surfaces and the two-dimensional contour plots were drawn using Design Expert 10
software (StatEase, USA) to investigate the interaction effects of variables and to identify
the optimal region for the proposed process.

3. Results and Discussion


3.1. Regression Model of Response
Regression analysis and estimation of regression coefficient were performed based on data
obtained by the BBD matrix. The matrix for nitrate removal with observed and predicted
response(s) for each set of reaction parameters is given in Table 3. The centre point (coded
values: 0, 0, 0) was repeated five times (Table 3: exp. no. 2, 3, 6, 8, 10) to evaluate the pure

Table 3. BBD matrix with observed and predicted nitrate removal (%) efficiency

Exp. no. pH Electrolysis time (min) Current (A) Yexp (%) Ypre (%)
1 9.5 100 1.5 85.82 87.69
2 9.5 60 1.0 83.14 86.23
3 9.5 60 1.0 84.55 86.23
4 9.5 20 1.5 56.06 60.51
5 12 20 1.0 49.57 67.49
6 9.5 60 1.0 82.02 86.23
7 9.5 100 0.5 70.03 66.25
8 9.5 60 1.0 81.01 86.23
9 12 60 1.5 80.47 76.73
10 9.5 60 1.0 80.00 86.23
11 12 60 0.5 57.59 61.39
12 7.0 60 1.5 79.47 75.81
13 9.5 20 0.5 54.32 52.43
14 7.0 100 1.0 87.96 87.72
15 7.0 60 0.5 64.79 68.10
16 12 100 1.0 78.02 74.61
17 7.0 20 1.0 53.58 53.77

INDIAN CHEMICAL ENGINEER 2017


8 ACHARYA ET AL .

error. Less than 2% deviation was observed in all the five identical runs, which confirmed the
reproducibility of experimental data.
An empirical relationship between % nitrate removal efficiency and the influencing process
variables was obtained by correlating experimental results with response functions using
Design expert 10 (StatEase, USA) regression program. The quadratic model describing the
response function with regression coefficients for nitrate removal from synthetic solution is
given in the following equation:

Y = +86.23 + (0.17∗A) + (10.25∗B) + (7.38∗C) − (6.69∗A∗B) + (0.29∗A∗C)


Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

+ (3.34∗B∗C) − (6.59∗A∗A) − (8.76∗B∗B) − (10.75∗C∗C), (7)

where %Y = removal efficiency; A = pH; B = electrolysis time (min); C = current (A).


It was inferred from Equation (7) and Table 3 that nitrate removal efficiency (%Y) from
synthetic solution can be fitted well to the developed quadratic polynomial equation.
Maximum 5% deviation was observed between experimental and predicted nitrate removal effi-
ciency. Electrolysis time and current exert a positive linear effect on removal of nitrate, whereas
the effect of pH on removal of nitrate was found to be less significant than other process par-
ameters. Mutual interaction effect of process parameters also played a dominant role in nitrate
removal. Negative coefficient was observed for the mutual effect of pH and electrolysis time,
whereas the effect of electrolysis time and current was depicted with the positive value of the
coefficient. These observations rationalized electrolysis time and current as the favourable par-
ameter for the removal of nitrate from aqueous solutions.

3.2. ANOVA Results


The ANOVA was conducted for nitrate removal and results are presented in Table 4.
ANOVA compares variation sources with Fisher distribution (F-test) to validate the feasi-
bility of the regression model. High F-value of 17.10 (Table 4) implied the acceptability of
the variation around its mean values. It clearly indicated the predictability of the
regression model at the 95% confidence interval. t-Test was applied to investigate the sig-
nificance of each coefficient of the developed regression model (shown in Equation (7))
and corresponding p-values are listed in Table 4. The probability of error value
(p-value) is used to determine whether the association between the response and each
term in the model is statistically significant. The p-value is defined as the smallest level
of significance leading to rejection of the null hypothesis and the interaction effects are
statistically significant when p < .05 [54]. The ANOVA results were found to be in agree-
ment with the inferences of Equation (7) where electrolysis time (B), current (C ), AC,
A 2, B 2, C 2 are significant model terms. However, a higher p-value of pH (A) and
mutual interaction effect of parameter AB (pH and time) may be related to the fact
that pH did not have much significant effect on removal efficiency. Similar results were
observed in 3D plots and the regression model for pH and reaction time. A saddle
region was obtained in the 3D plot of pH vs. time, which suggested that mutual inter-
action effect could negatively affect nitrate removal efficiency beyond a certain limit of
both parameters. It was also inferred from the regression model polynomial equation
INDIAN CHEMICAL ENGINEER 2017
Statistical Optimization of Electrocoagulation Process 9

Table 4. ANOVA analysis for the quadratic model developed for synthetic nitrate solution

Source Sum of squares Df Mean square F-value p-Value

Model 2603.81 9 289.31 17.10 0.0006 Significant


A 0.23 1 0.23 0.014 0.9102
B 839.68 1 839.68 49.62 0.0002
C 436.01 1 436.01 25.77 0.0014
AB 179.16 1 179.16 10.59 0.0140
AC 0.33 1 0.33 0.020 0.8928
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

BC 44.69 1 44.69 2.64 0.1482


A2 183.06 1 183.06 10.82 0.0133
B2 323.01 1 323.01 19.09 0.0033
C2 486.92 1 486.92 28.78 0.0010
Residual 118.45 7 16.92
Lack of fit 85.36 3 28.45 3.44 0.1319 Not significant
Pure error 33.09 4 8.27
Cor total 2722.26 16
Std. dev. 4.11 R2 0.9565
Mean 73.94 Adj. R 2 0.9005
*C.V% 5.56 Adeq. precision 11.174
*PRESS 1417.49

*C.V. %: Coefficient of variation, PRESS : Predicted residual sum of squares

where a high value of negative coefficient was achieved for the mutual interaction effect of
reaction pH and reaction time. p-Value (0.1482) for the mutual interaction effect of par-
ameter BC (time and current) was also found to be greater than 0.05. However, values
are not considerably high which can be explained on the basis of relative error in data col-
lection [50]. Parameters having p-values> .05 can be eliminated using the stepwise elimin-
ation method. A low value of 5.56 of coefficient of variation (C.V.) indicated a higher
degree of precision and reliability of the experiments.
A high value of the regression coefficient (R 2 = .95) and adjusted regression coefficient
(R2adj = 0.9) corroborated the suitability of the model. The closeness of these values augmented
the validation of the model. It was observed that all values are within the ± 4% of the straight
line, which indicates good acceptability of predicted nitrate removal efficiency using the
regression model. ANOVA results are presented in Fig. 2.

3.3. Optimization of Process Parameters


3.3.1. Response Surface Plots
Behaviour and extent of the interacting parameters on the efficiency depend on the shape of the
contours. Elliptical nature of the 2D contour plots shows the good interaction effect, whereas
straight lines demonstrate poor interaction between process parameters. 2D contours and 3D
interaction plots are shown in Figs. 3–5.
INDIAN CHEMICAL ENGINEER 2017
10 ACHARYA ET AL .
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

Fig. 2. Graphical representation of observed vs. predicted nitrate removal efficiency.

Fig. 3(A, B) demonstrates the interaction effect of electrolysis time and reaction pH in the
form of 3D plots and 2D contours, respectively. It can be depicted from 3D plot and 2D con-
tours that initially, efficiency increases with the increase in pH and increase in electrolysis time.
However, a saddle region was observed at pH 10, which indicated the negative effect of reaction
pH beyond a certain limit on nitrate removal efficiency. It can be explained by the fact that the
oxide surface has a net negative charge at alkaline pH, and would tend to repulse the anionic
nitrate in solution [47].
Fig. 4(A, B) demonstrates 3D plots and 2D contours, respectively, to investigate the effect
of reaction pH and current on nitrate removal efficiency. A significant mutual interaction effect
between these two parameters was observed. Nearly, 78% removal of nitrate was observed at
high level of current and low level of pH values. This efficiency is approximately 30% higher

Fig. 3. (A) 3D plots to show the interaction between electrolysis time and reaction pH. (B): 2D contours
to show the interaction between electrolysis time and reaction pH.

INDIAN CHEMICAL ENGINEER 2017


Statistical Optimization of Electrocoagulation Process 11
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

Fig. 4. (A): 3D plots to show the interaction between reaction pH and current. (B): 2D contours to show
the interaction between reaction pH and current.

Fig. 5. (A): 3D plots to show the interaction between electrolysis time and current. (B): 2D contours to
show the interaction between electrolysis time and current.

than the efficiency observed at low level of both the parameters. Nearly, 10% increment was
observed when pH was increased from its low level to high level, which clearly indicated that
the current played a dominant role than pH in the removal of nitrate using the EC process. Con-
centric elliptical 2D contours are visible in Fig. 4(B), which substantiate the inferences derived
from Equation (7). A positive coefficient (0.29) was obtained in the regression model of
response to establish a mutual interaction between pH and current for the effective removal
of nitrate from synthetic solution.
The effect of electrolysis time and current is demonstrated in Fig. 5(A, B) with the help of
3D plots and 2D Contours. It can be depicted from Fig. 5(A) that both factors play a significant

INDIAN CHEMICAL ENGINEER 2017


12 ACHARYA ET AL .

role on nitrate removal efficiency. More than 35% increment in the removal of nitrate was
observed when both factors were moved from low level to high level. Results were found to
be in concordance with the literature. Vasudevan et al. [46] suggested a significant increase
in the coagulation rate when the current is increased; thus, the removal of nitrate depends
upon the availability of binding sites.
From the above observations, it was concluded that the removal of nitrate from aqueous
solution using the EC process depends primarily on electrolysis time and current, whereas reac-
tion pH does not show a remarkable change in removal efficiency of nitrate. The optimized
reaction conditions for the proposed EC process were electrolysis time = 77 min, current =
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

1.2 A and solution pH = 9.8.

3.3.2. Model Validation


The major objective of the present study was to maximize the removal of nitrate from aqueous
stream with recalculation of all responsible process parameters using desirability functions.
Point prediction feature of the Design Expert 10 software was used to get different combi-
nations of process parameters for maximizing the target response. Confirmatory experiments
were conducted in triplicate at the nearest possible values of process parameters to validate
the results obtained from numerical optimization. Table 5 illustrates the results obtained for
validating the regression model. Maximum ± 4% deviation was observed in theoretical and
experimental data. Experiments were also performed on real groundwater using optimum con-
ditions obtained from theoretical investigation and results obtained were in agreement with the
developed model. This indicates that BBD in conjunction with desirability functions can be effi-
caciously applied to optimize the design of experiments for removal of nitrate using the EC
process.

4. Conclusion
The present work is a pioneer effort for statistical optimization of the removal of nitrates from
aqueous solutions using the EC process. The BBD matrix in conjunction with RSM was
employed for statistical optimization of the proposed process. A quadratic regression model
was developed with regression coefficient more than 0.95, which indicated high correlation
between observed and predicted model. ANOVA studies and 3D plots suggested electrolysis
time and current as the dominant process parameters for nitrate removal. Electrolysis time
= 77 min, current = 1.2 A and solution pH = 9.8 were obtained as the optimum reaction

Table 5. Validation experiments.

Current
Exp. No. pH Electrolysis time (min) (amp) Yexp (%) Ypre (%) % deviation

18 10 75 1.2 87.62 89.34 1.963022


19 10 77 1.2 87.74 89.47 1.971735
20 10 80 1.5 89.55 86.38 −3.53992

INDIAN CHEMICAL ENGINEER 2017


Statistical Optimization of Electrocoagulation Process 13

conditions to achieve nearly 89.4% removal of nitrate from water samples. The present study
clearly suggests that RSM is an effective tool for predicting interaction of influencing process par-
ameters and for the large-scale optimization of the EC process at industrial platform.

Disclosure statement
Authors have no potential conflict of interest.
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

References
[1] Nancharaiah, Y.V. and Venugopalan, P.V., “Denitrification of Synthetic Concentrated Nitrate
Wastes by Aerobic Granular Sludge Under Anoxic Conditions”, Chemosphere, 85, pp. 683–688
(2011).
[2] Koster, J.R., Cardenas, L., Senbayrama, M., Bol, R., Well, R., Butler, M., Muhlinga, K.H. and
Dittert, K., “Rapid Shift From Denitrification to Nitrification in Soil After Biogas Residue
Application as Indicated by Nitrous Oxide Isotopomers”, Soil Biol. Biochem., 43, pp. 1671–1677
(2011).
[3] Ansari, M.H. and Parsa, J.B., “Removal of Nitrate From Water by Conducting Polyaniline via
Electrically Switching ion Exchange Method in a Dual Cell Reactor: Optimizing and
Modeling”, Sep. Purif. Technol., 169, pp. 158–170 (2016).
[4] Ghafari, S., Hasan, M. and Aroua, M.K., “Bio-electrochemical Removal of Nitrate From Water
and Wastewater—A Review”, Bioresour. Tech., 99, pp. 3965–3974 (2008).
[5] WHO, Nitrate and Nitrite in Drinking Water, WHO/SDE/WSH/07.01/16/Rev/1,WHO Press,
Geneva, Switzerland (2011).
[6] BIS (Bureau of Indian Standards), Specification for drinking water. IS: 10500, New Delhi, India
(2012).
[7] Rossi, F., Motta, O., Matrella, S., Proto, A. and Vigliotta, G., “Nitrate Removal From Wastewater
Through Biological Denitrification with OGA 24 in a Batch Reactor”, Water, 7, pp. 51–62 (2015).
[8] Cheikh, A., Yala, A., Drouiche, N., Abdi, N., Lounici, H. and Mameri, N., “Denitrification of
Water in Packed Beds Using Bacterial Biomass Immobilized on Waste Plastics as Supports”,
Ecol. Eng., 53, pp. 329–334 (2013).
[9] De Filippis, P., Di Palma, L., Scarsella, M. and Verdone, N., “Biological Denitrification of High
Rate Wastewater: A Comparison Between Three Electron Donors”, Chem. Eng. Trans., 32, pp.
319–324 (2013).
[10] Dholam Sujit, S., Attarde, S.B., Wagh, N.D. and Ingale, S.T., “De-nitrification of Wastewater Using
Sludge and Methanol as a Carbon Source”, Univ. J. Environ. Res. Technol., 4, pp. 172–177 (2014).
[11] Khanitchaidecha, T. Sumino and Kazama, F., “Influence of Carbon Source on Biological Nitrogen
Removal by Immobilised Bacteria”, J. Water. Res. Prot., 2, pp. 527–531 (2010).
[12] Park, Y. and Yoo, Y.J. “Biological Nitrate Removal in Industrial Wastewater Treatment: Which
Electron Donor we can Choose”, Appl. Microbial Biotechnol., 82, pp. 415–429 (2009).
[13] Boumediene, M. and Achour, D., “Denitrification of the Underground Waters by Specific Resin
Exchange of Ion”, Desalination, 168, pp. 187–194 (2004).
[14] Pintor, A., Batista, J. and Levec, J., “Integrated ion Exchange/Catalytic Process for Efficient
Removal of Nitrates From Drinking Water”, Chem. Eng. Sci., 56, pp. 1551–1559 (2001).

INDIAN CHEMICAL ENGINEER 2017


14 ACHARYA ET AL .

[15] Schoeman, J.J. and Steyn, A., “Nitrate Removal with Reverse Osmosis in a Rural Area in South
Africa”, Desalination, 155, pp. 15–26 (2003).
[16] Yorn, T., Shon, Z.H. and Lee, G., “Parametric Studies on the Performance of Anion Exchange for
Nitrate Removal”, Korean. J. Chem. Eng., 18, pp. 170–177 (2001).
[17] Samatya, S., Kabay, N., Yuksel, U., Arda, M. and Yuksel, M., “Removal of Nitrate From Aqueous
Solution by Nitrate Selective ion Exchange Resins”, React. Funct. Polym., 66, pp. 1206–1214 (2006).
[18] Epsztein, R., Nir, O., Lahav, O. and Green, M., “Selective Nitrate Removal From Groundwater
Using a Hybrid Nanofiltration–Reverse Osmosis Filtration Scheme”, Chem. Eng. J., 279, pp.
372–378 (2015).
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

[19] Ahn, J.H., Choo, K.H. and Park, H.S., “Reverse Osmosis Membrane Treatment of Acidic Etchant
Wastewater: Effect of Neutralization and Polyelectrolyte Coating on Nitrate Removal”, J. Memb.
Sci., 310, pp. 296–302 (2008).
[20] Richards, L.A., Vuachere, M. and Schafer, A.I., “Impact of pH on the Removal of Fluoride,
Nitrate and Boron by Nanofiltration/Reverse Osmosis”, Desalination, 261, pp. 331–337 (2010).
[21] Cheikh, A., Grib, H., Drouiche, N., Abdi, N., Lounici, H. and Mameri, N., “Water Denitrification
by a Hybrid Process Combining a new Bioreactor and Conventional Electrodialysis”, Chem. Eng.
Process., 63, pp. 1–6 (2013).
[22] McAdam, E.J. and Judd, S.J., “A Review of Membrane Bioreactor Potential for Nitrate Removal
From Drinking Water”, Desalination, 196, pp. 135–148 (2006).
[23] Devadas, A., Vasudevan, S. and Epron, F., “Nitrate Reduction in Water: Influence of the Addition
of a Second Metal on the Performances of the Pd/CeO2 Catalyst”, J. Hazard. Mater., 185, pp.
1412–1417 (2011).
[24] Chollier-Brym, M.J., Gavagnin, R., Strukul, G., Marella, M., Tomaselli, M. and Ruiz, P., “New Insight
in the Solid State Characteristics, in the Possible Intermediates and on the Reactivity of Pd–Cu and Pd–
Sn Catalysts, Used in Denitratation of Drinking Water”, Catal. Today, 75, pp. 49–55 (2002).
[25] Ganesan, P., Kamaraj, R. and Vasudevan, S., “Application of Isotherm, Kinetic and
Thermodynamic Models for the Adsorption of Nitrate Ions on Graphene From Aqueous
Solution”, J. Taiwan Inst. Chem. Eng., 44, pp. 808–814 (2013).
[26] Mizuta, K., Matsumoto, T., Hatate, Y., Nishihara, K. and Nakanishi, T., “Removal of Nitrate-
Nitrogen From Drinking Water Using Bamboo Powder Charcoal”, Bioresour. Technol., 95, pp.
255–257 (2004).
[27] Mena-Duran, C.J., Sun Kou, M.R., Lopez, T., Azamar-Barrios, J.A., Aguilar, D.H., Domínguez,
M.I., Odriozola, J.A. and Quintana, P., “Nitrate Removal Using Natural Clays Modified by Acid
Thermoactivation”, Appl. Surf. Sci., 253, pp. 5762–5766 (2007).
[28] Acharya, S. and Sharma, S.K., “Groundwater Assessment and its Electrochemical Treatment”,
Int. J. Adv. Technol. Eng. Sci., 4, pp. 21–30 (2016).
[29] Khandegar, V. and Saroha, A.K., “Electrocoagulation for the Treatment of textile Industry
Effluent – A Review”, J. Environ. Manage., 128, pp. 949–963 (2013).
[30] Yilmaz, A.E., Boncukcuoğlu, R. and Kocakerim, M.M., “A Quantitative Comparison Between
Electrocoagulation and Chemical Coagulation for Boron Removal From Boron-Containing
Solution”, J. Hazard. Mater., 149, pp. 475–481 (2007).
[31] Moussavi, G., Khosravi, R. and Farzadkia, M., “Removal of Petroleum Hydrocarbons From
Contaminated Groundwater Using an Electrocoagulation Process: Batch and Continuous
Experiments”, Desalination, 278, pp. 288–294 (2011).

INDIAN CHEMICAL ENGINEER 2017


Statistical Optimization of Electrocoagulation Process 15

[32] Zongo, I., Hama Maiga, A., Wéthé, J., Valentin, G., Leclerc, J.P., Paternotte, G. and Lapicque, F.,
“Electrocoagulation for the Treatment of textile Wastewaters with Al or Fe Electrodes: Compared
Variations of COD Levels, Turbidity and Absorbance”, J. Hazard. Mater., 169, pp. 70–776 (2009).
[33] Khandegar, V. and Saroha, A.K., “Electrochemical Treatment of Textile Effluent Containing Acid
Red 131 Dye”, J. Hazard. Toxic. Radioact. Waste., 18, pp. 38–44 (2014).
[34] Lemlikchi, W., Khaldi, S., Mecherri, M.O., Lounici, H. and Drouiche, N., “Degradation of Disperse
Red 167 Azo Dye by Bipolar Electrocoagulation”, Sep. Sci. Technol., 47, pp. 1682–1688 (2012).
[35] Farhadi, S., Aminzadeh, B., Torabian, A., Khatibikamal, V. and Alizadeh Fard, M, “Comparison
of COD Removal From Pharmaceutical Wastewater by Electrocoagulation,
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

Photoelectrocoagulation, Peroxi-Electrocoagulation and Peroxi-Photoelectrocoagulation


Processes”, J. Hazard. Mater., 219–220, pp. 35–42 (2012).
[36] Verma, S.K., Khandegar, V. and Saroha, A.K., “Removal of Chromium From Electroplating Industry
Effluent Using Electrocoagulation”, J. Hazard. Toxic. Radio. Waste., 17, pp. 146–152 (2013).
[37] Kamaraj, R. and Vasudevan, S., “Facile one-pot Electrosynthesis of Al(OH)3 – Kinetics and
Equilibrium Modeling for Adsorption of 2,4,5-Trichlorophenoxyacetic Acid From Aqueous
Solution”, New. J. Chem., 40, pp. 2249–2258 (2016).
[38] Kamaraj, R., Davidson, D.J., Sozhan, G. and Vasudevan, S., “Adsorption of Herbicide 2-(2,4-
Dichlorophenoxy)Propanoic Acid by Electrochemically Generated Aluminum Hydroxides: an
Alternative to Chemical Dosing”, RSC Adv., 5, pp. 39799–39809 (2015).
[39] Heffron, J., Marhefke, M. and Mayer, B.K., “Removal of Trace Metal Contaminants from Potable
Water by Electrocoagulation”, Sci. Rep., 6, pp. 28478–28487 (2016).
[40] Cheballah, K., Sahmoune, A., Messaoudi, K., Drouiche, N. and Lounici, H., “Simultaneous
Removal of Hexavalent Chromium and COD From Industrial Wastewater by Bipolar
Electrocoagulation”, Chem. Eng. Process., 96, pp. 94–99 (2015).
[41] Zeboudji, B., Drouiche, N., Lounici, H., Mameri, N., and Ghaffour, N., “The Influence of
Parameters Affecting Boron Removal by Electrocoagulation Process”, Sep. Sci. Technol., 48, pp.
1280–1288 (2013).
[42] Kamaraj, R. and Vasudevan, S., “Facile one-pot Synthesis of Nano-Zinc Hydroxide by Electro-
Dissolution of Zinc as a Sacrificial Anode and the Application for Adsorption of Th4+, U4+,
and Ce4+ From Aqueous Solution”, Res. Chem. Intermed., 42, pp. 4077–4095 (2016).
[43] Palahouane, B., Drouiche, N., Aoudj, S., Bensadok, K., “Cost-effective Electrocoagulation Process
for the Remediation of Fluoride From Pretreated Photovoltaic Wastewater”, J. Ind. Eng. Chem., 22,
pp. 127–131 (2015).
[44] Drouiche, N., Aoudj, S., Lounici, H., Drouiche, M., Ouslimane, T. and Ghaffour, N., “Fluoride
Removal From Pretreated Photovoltaic Wastewater by Electrocoagulation: An Investigation of
The Effect of Operational Parameters”, Procedia. Eng., 33, pp. 385–391 (2012).
[45] Pak, K.S., “Factors Influencing Treatment of Nitrate Contaminated Water using Batch
Electrocoagulation Process”, Int. J. Curr. Eng. Technol., 5, pp. 714–718 (2015).
[46] Vasudevan, S., Epron, F., Lakshmi, J., Ravichandran, S., Mohan, S. and Sozhan, G., “Removal of
NO3 – From Drinking Water by Electrocoagulation – An Alternate Approach”, Clean. Soil. Air.
Water., 38, pp. 225–229 (2010).
[47] Kamaraj, R., Pandiarajan, A., Jayakiruba, S., Naushad, M. and Vasudevan, S., “Kinetics,
Thermodynamics and Isotherm Modeling for Removal of Nitrate From Liquids by facile one-
pot Electrosynthesized Nano Zinc Hydroxide”, J. Mol. Liquids., 215, pp. 204–211 (2016).

INDIAN CHEMICAL ENGINEER 2017


16 ACHARYA ET AL .

[48] Kumar, N.S. and Goel, S., “Factors Influencing Arsenic and Nitrate Removal From Drinking
Water in a Continuous Flow Electrocoagulation (EC) Process”, J. Hazard. Mater., 173, pp. 528–
533 (2010).
[49] Lakshmi, J., Sozhan, G. and Vasudevan, S., “Recovery of Hydrogen and Removal of Nitrate From
Water by Electrocoagulation Process”, Environ. Sci. Pollut. Res., 20, pp. 2184–2192 (2013).
[50] Chauhan, G., Pant, K.K. and Nigam, K.D.P., “Development of Green Technology for Extraction
of Nickel from Spent Catalyst and its Optimization Using Response Surface Methodology”, Green.
Proc. Synth., 2, pp. 259–271 (2013).
[51] Behloul, M., Grib, H., Drouiche, N., Abdi, N., Lounici, H. and Mameri, N., “Removal of
Downloaded by [GGS Indraprastha University] at 22:29 07 September 2017

Malathion Pesticide From Polluted Solutions by Electrocoagulation: Modeling of Experimental


Results Using Response Surface Methodology”, Sep. Sci. Technol., 48, pp. 664–672 (2013).
[52] Chenna, M., Messaoudi, K., Drouiche, N. and Lounici, H., “Study and Modeling of the
Organophosphorus Compound Degradation by Photolysis of Hydrogen Peroxide in Aqueous
media by Using Experimental Response Surface Design”, J. Ind. Eng. Chem., 33, pp. 307–315
(2016).
[53] Demim, S., Drouiche, N., Aouabed, A., Benayad, T., Dendene-Badache, O. and Semsari, S.,
“Cadmium and Nickel: Assessment of the Physiological Effects and Heavy Metal Removal
Using a Response Surface Approach by L. gibba”, Ecol. Eng., 61, pp. 426–435 (2013).
[54] Demim, S., Drouiche, N., Aouabed, A., Benayad, T., Couderchet, M. and Semsari, S., “Study of
Heavy Metal Removal From Heavy Metal Mixture Using the CCD Method”, J. Ind. Eng.
Chem., 20, pp. 512–520 (2014).
[55] Ghanim, A.N., “Application of Response Surface Methodology to Optimize Nitrate Removal from
Wastewater by Electrocoagulation”, Int. J. Scient. Eng. Res., 4, pp. 1410–1416 (2013).
[56] Hossini, H. and Rezaee, A., “Optimization of Nitrate Reduction by Electrocoagulation Using
Response Surface Methodology”, Health. Scope, 3, pp. 1–6 (2014).
[57] Bhanu, P.U. and Prasad, B., “Removal of Bromide and Nitrate from Petroleum Refinery
Wastewater Using Electrocoagulation”, ICARMMIEM-2014, pp. 26–29 (2014).
[58] Demim, S., Drouiche, N., Aouabed, A. and Semsari, S., “CCD Study on the Ecophysiological
Effects of Heavy Metals on Lemna gibba”, Ecol. Eng., 57, pp. 302–313 (2013).
[59] Emamjomeh, M. and Sivakumar, M., “Denitrification Using a Monopolar Electrocoagulation/
Flotation (ECF) Process”, J. Environ. Manage., 91, pp. 516–522 (2009).
[60] Almeida, J.S., Reis, M.A.M. and Carrondo, M.J.T., “Competition Between Nitrate and
Nitrite Reduction in Denitrification ByPseudomonas fluorescens”, Biotechnol. Bioeng., 46,
pp. 476–484 (1995).
[61] Zayed, G. and Winter, J., “Removal of Organic Pollutants and of Nitrate From Wastewater From
the Dairy Industry by Denitrification”, Appl. Microbiol. Biotechnol., 49, pp. 469–474 (1998).
[62] Peyton, B.M., Mormile, M.R. and Petersen, J.N., “Nitrate Reduction with Halomonas campisalis:
Kinetics of Denitrification at pH 9 and 12.5% NaCl”, Water Res., 35, pp. 4237–4242 (2001).
[63] Drouiche, N., Aoudj, S., Hecini, M., Ghaffour, N., Lounici, H. and Mameri, N., “Study on the
Treatment of Photovoltaic Wastewater Using Electrocoagulation: Fluoride Removal with
Aluminium Electrodes—Characteristics of Products”, J. Hazard. Mater., 169, pp. 65–69 (2009).
[64] Li, M., Feng, C., Zhang, Z., Chen, R., Xue, Q., Gao, C. and Sugiura, N., “Optimization of Process
Parameters for Electrochemical Nitrate Removal Using Box–Behnken Design”, Electrochim. Acta,
56, pp. 265–270 (2010).

INDIAN CHEMICAL ENGINEER 2017

You might also like