You are on page 1of 16

Environmental Technology & Innovation 22 (2021) 101448

Contents lists available at ScienceDirect

Environmental Technology & Innovation


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

Continuous flowing electrocoagulation reactor for efficient


removal of azo dyes: Kinetic and isotherm studies of
adsorption

Ze Wu, Jie Dong, Yingwu Yao, Yang Yang, Feng Wei
Hebei University of Technology, School of Chemical Engineering and Technology, Tianjin 300130, PR China

article info a b s t r a c t

Article history: The electrocoagulation process is an effective and economical technology for the treat-
Received 2 November 2020 ment of industrial wastewater. So, it is particularly important to develop an efficient
Received in revised form 15 February 2021 electrocoagulation reactor. To this end, a novel electrocoagulation reactor is described
Accepted 19 February 2021
to effectively remove azo dyes from aqueous solutions utilizing electrocoagulation, in
Available online 23 February 2021
which the electrode plate is bent. Continuous wastewater flows in the reactor will
Keywords: promote faster and more stable growth of flocs This paper took methyl orange as
Continuous electrocoagulation reactor a study object. It was concluded that methyl orange decolorization rate was 92.35%
Azo dye under the optimal conditions while flat electrode under the same conditions of methyl
Response surface methodology orange decolorization rate was only 58.9%. The results meant that the second-order
Isotherm kinetic model was most in line with the experimental results, which indicated that
Kinetic the chemisorption mechanism controls the adsorption of methyl orange. Furthermore,
the results showed that the Langmuir isotherm model was more in line with the
experimental results, and showed that the experiment was single-layer adsorption on
the adsorbent surface. Under the optimal conditions of the electrocoagulation process,
electrode consumption is 0.2376 g, and the electrical energy consumption is 1.5925 KWh
m−3 during a single operation.
© 2021 Elsevier B.V. All rights reserved.

1. Introduction

As we all know, water is the source of life, as one of the basic needs of living organisms has played an irreplaceable
role, and industrial production is inseparable from water, but water shortage and depletion of freshwater resources is a
major environmental challenge facing mankind (Hogeboom, 2020). Moreover, the discharge of sewage is increasing year
by year (Zhang et al., 2015). Industrial wastewater produced by the textile industry is one of the most serious sources of
environmental pollution, it is composition is complex, including organic and inorganic pollutants, and contains various
types of dyes (Samsami et al., 2020), According to the chromophore structure, dyes can be divided into acid, basic, disperse,
active, azo, diazo and anthraquinone (Su et al., 2016). Dye molecules cause mutations in the human organs, brain, central
nervous system, and reproductive system, which are carcinogenic and cause dysfunction (Zhou et al., 2019). More than
70% of the approximately 900,000 metric tons of dyes produced each year belong to the azo group (Rawat et al., 2016).
Azo dyes cause dyes to do great harm to human health and the environment (Rawat et al., 2016; Tkaczyk et al., 2020).

∗ Corresponding author.
E-mail address: weifeng@hebut.edu.cn (F. Wei).

https://doi.org/10.1016/j.eti.2021.101448
2352-1864/© 2021 Elsevier B.V. All rights reserved.
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Biological and chemical methods are two main methods for treating dye wastewater. Biodegradation has treatment
methods such as activated sludge (Mirbolooki et al., 2017), anaerobic sludge (Kong et al., 2019), aerobic (de Sousa Rollem-
berg et al., 2018). Due to their process characteristics, they always took up a large amount of physical space and had long
processing cycles (Brillas and Martínez-Huitle, 2015; Nunez et al., 2019), so sometimes it is not very convenient. Chemical
treatments such as adsorption (Dolatabadi et al., 2020; Feng et al., 2020), coagulation (Kuppusamy et al., 2017), Fenton
(Ozbey Unal et al., 2019), photocatalysis (Nguyen and Juang, 2019), etc, were generally highly efficient, but they need to
add chemical reagents leads to higher costs. Electrochemistry has been widely used in chemical, metallurgical, mechanical,
light industry, medicine, materials, energy, metal corrosion and protection, environmental science and other scientific and
technological fields, electrochemical treatment of wastewater methods have electric Fenton (Malakootian and Moridi,
2017), electrocatalysis (Suhadolnik et al., 2019), electrocoagulation (EC) (Ozyonar et al., 2020). EC is a fully studied and
valid precipitation method, which can remove a variety of industrial pollutants. This process forms metal hydroxides and
oxyhydroxides by electrolytic oxidation at the ‘‘sacrificial anode’’, and then these types of coagulants offer active surfaces
to destroy and adsorption of contaminants (Fan et al., 2020). In addition to the above contaminant removal methods,
electrochemical methods were applied to the determination of some wastewater parameters (Ahmadzadeh et al., 2015,
2019; Avazpour et al., 2020; Badakhshan et al., 2019). The complexation reaction of new reagents on organic ions and
metal ions were studied in references (Ahmadzadeh et al., 2011; Rezayi et al., 2011a,b; Rounaghi et al., 2009) in order to
develop the new electrodes or adsorption materials.
EC can address a variety of industrial wastewater, such as wastewater containing chromium, zinc, copper, and
other heavy metals (Kim et al., 2020), textile industry wastewater (Bener et al., 2019), paper industry wastewater
(Barhoumi et al., 2019), pharmaceutical wastewater (Zaied et al., 2020), etc., with very good effect. Studies have explored
electrode linking methods, including Monopole parallel (MP-P), Monopole series (MP-S), and Bipolar series (BP-S). In
a comprehensive consideration, the MP-P configuration was the most cost-effective (Nasrullah et al., 2018). There was
also a rotating anode electrode reactor that can better solve the problem of electrode passivation, the dead zone inside
the reactor (Al-Raad et al., 2020; Villalobos-Lara et al., 2020). Internal loop split-plate airlift reactor reduced the energy
consumption (EEC) of wastewater treated in refineries (Ammar et al., 2019). The electrocoagulation membrane reactor
mainly filtered the wastewater after EC and did not improve the efficiency of the reactor (Xu et al., 2021). In addition,
there was a combination of a photovoltaic system and electrocoagulation reactor (Combatt et al., 2020), a fixed bed
anode electrocoagulation reactor (Rodrigues et al., 2020), etc. However, most EC studies were conducted utilizing small
intermittent reactors, while most industrial applications are based on continuous flow patterns. In addition, there were
some researches on the adsorption of heavy metals and less research on the process of dye adsorption (Moussa et al.,
2017).
In this study, a continuous electrocoagulation reactor was designed to degrade azo dyes by electrocoagulation with
a folded plate electrode. This research used the single-factor experiment to analyze the effect of each parameter on the
experiment. Response surface methodology (RSM) was used to study the main effects of the initial concentration of methyl
orange (MO), the current density, and the reaction time. The reliability of the model was tested by analysis of variance
(ANOVA). In addition, pseudo-first-order and pseudo-second-order kinetic models were studied to find out the exact
process of using EC processes to remove azo. The adsorption of contaminants on the surface of the resulting adsorbent
was studied utilizing two common adsorption iso-temperature lines, Freundlich and Langmuir isotherm models. Finally,
the experimental electrode consumption (ELC) and energy consumption were calculated.

2. Devices and methods

2.1. Experimental setup and procedure

The electrocoagulation reactor is made of perspex, and the dimensions are 115 mm × 100 mm × 100 mm. As shown
in Fig. 1, the device is made of folded plate iron sheets (see Fig. 2) as cathode and anode. The iron sheet material is
Q235, and the purity is more than 97%. Hydraulic plate bender is used for bending. The electrodes are washed by H2 SO4
before use, rinsed with distilled water, and dried. The synthetic wastewater is injected into the reactor with 1000 ml and
a liquid level of 8.7 cm. The current density of 10 mA cm−2 is provided using a precision digital DC power supply (Wei
et al., 2012). The experimental wastewater is recycled with a water pump that provides a flow rate of 0–105 L h−1 . All
experiments are performed in batch mode and at room temperature.

2.2. Experimental design

2.2.1. Single-factor experimental design


In this study, the influence of the flow rate in the reactor, plate spacing, and plate folding angle on the experiment
were studied to explore the optimal variable parameters.
Table 1 provides the range and level of variables in actual units.
2
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Fig. 1. Electrocoagulation reactor.

Fig. 2. Fold plate electrode.

Table 1
The range and level of variables.
Variables Range
Flow rate (L h−1 ) 15 45 60 75 105
Polar plate spacing (cm) 1 1.5 2 2.5 3
Angle(◦ ) 45 60 75 90 105

2.2.2. RSM experimental design


The EC process can improve process efficiency by optimizing the process and operating variables. Experimental data
was addressed using the Design-Expert software to optimize operational variables. The response surface methodology
was established using Box–Benkhen Experimental Design (BBD) technology, and three important working parameters of
initial methyl orange solution concentration, current density, and reaction time, were studied and optimized at three
levels through ANOVA, mathematical modeling, and response surface (Ba Mohammed et al., 2020; Khorram and Fallah,
2018). The regression model is built to:
3 3 3 3
∑ ∑ ∑ ∑
Y = a0 + ai X i + aii Xi2 + aij Xi Xj (1)
i=1 i=1 i=1 j=i+1

In this equation, Y is the response, a0 is the constant, ai is the linear coefficient, aii is the secondary coefficient, aij is
the interaction coefficient, Xi and Xj are independent variables.
The fitting excellence test (R2 test) is selected to evaluate the significance of the model. The determination coefficient
R2 can indicate the degree of difference between the response value and the true value, which is between 0–1 value. The
closer R2 is to 1, the smaller the difference between the response value and the true value. When R2 is 1, it means that
3
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Table 2
The code and level of factors.
The encoding value Methyl orange concentration (mg/L) Current density (ma) Reaction time (min)
−1 100 5 30
0 200 10 40
1 300 15 50

the two are the same.


n n
SR ∑ ∑
R2 = = (ŷi − y)2 / (yi − y)2 (2)
ST
i=1 i=1

In this: SR is the sum of regression squares; ST is the sum of total deviation squares; yi , ŷi and y are test values,
predictions, and test averages for the point response, respectively; n is the sample capacity.
Table 2 shows the code and level of factors in coding units.

2.3. Chemicals

Methyl Orange (C14 H14 N3 SO3 Na) was purchased from Chemart (Tianjin, China) Chemical Technology Company with
a purity of AR. The 0.01n H2 SO4 and NaOH solution and Na2 SO4 powder purity AR were purchased from Chemart. All
solutions were prepared with deionized water and at least analytical chemicals. Synthetic wastewater was made from
deionized water with electrolyte sodium sulfate and dissolved methyl orange, and PH was adjusted to 6 (Khandegar and
Saroha, 2013), measured with PHS-3C in Shanghai, China.

2.4. Analysis method

The experiment time of each EC is 50 min. Samples were taken every ten minutes. After the samples were filtered,
the corresponding maximum absorption wavelength was measured by INESAL5UV–VIS (Shanghai, China) and calculated
at 460 nm color ratio. The decoloring efficiency is computed by the following formula (Abbasi et al., 2020):
Ao − At
C olorremov al(%) = × 100 (3)
Ao
Where A0 and At respectively express the initial and final absorption of wastewater.
Calculate the energy consumption (KWh m−3 ) in the experiment by the formula:
UIT
EEC = (4)
V
Among them, V, I, t, and U express sample volume (L), applied current (A), reaction time (h), and applied voltage (V).

3. Results and discussion

The single-factor experiment explored the effects of flow, plate spacing, and plate angle on the experiment, with an
initial conditional current density of 10 mA cm−2 , an initial pH of 6, and an initial concentration of methyl orange of 100
mg L−1 . The RSM experiment explored the initial concentration of methyl orange, current density, electrolyte time of
individuals, and interactions. The results of the experiment were fitted with adsorption dynamics and iso-temperature
models, and the loss of the experiment was calculated.

3.1. The single-factor experiment

3.1.1. The effect of flow


The size of the processing volume reflects the processing performance of a reactor and in general the larger the
processing volume, the better. But it cannot be infinite. Proper flow can promote the growth of flocculants, accelerate
the rate of pollutant removal, and purify sewage as soon as possible. The group of the experiment was conducted under
the initial conditions of the edge length and degree of folding plate electrode of 3 cm and 90◦ , and the electrode spacing
of 2 cm. The results of wastewater treatment are shown in Fig. 3. When all flow is within 30 min, the decolorization
rate increases the fastest. When all flow is in the 30th minute, the decolorization rate reaches more than 80%. However,
when all flow is 30 min later, the decolorization rate increases slowly. Too large or too small flow is not conducive to
the improvement of decolorization rate. The decolorization rate shows a tendency to increase and then decrease with
the increase inflow. It can be seen from the figure that the decolorization rate corresponding to the flow rate of 60 L h−1
at the beginning of the experiment is higher than that under other flow. This indicates that EC collate cannot combine
with methyl orange particles better when the flow rate is less than 60 L h−1 . When the flow rate is more than 60 L h−1 ,
the collate is destroyed by high-speed flow, thus reducing the flocculation effect. When the flow rate is 60 L h−1 , the
experimental effect is the best.
4
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Fig. 3. UV–visible measurements to assess the effect of flow on the experiment.

Fig. 4. UV–visible measurements to assess the effect of plate distance on the experiment.

3.1.2. Effect of plate distance


In the practical application process, the size of the board spacing can affect the flow rate of waste liquid between the
plates. Because of the folding plate electrode, the flow between different plates will have different turbulence effects and
thus affect the efficiency of the experiment. Therefore, it is important to explore the spacing of the electrode plates. The
group of the experiment was conducted under the initial conditions of the folding plate electrodes with an edge length
of 3 cm and 90◦ and the flow rate of 60 L h−1 . The effect of plate spacing from 1 cm to 3 cm on electrocoagulation was
investigated. The results of wastewater treatment are shown in Fig. 4. It shows that the decolorization rate increases
linearly with time, and it reaches more than 90% in 30 min when the distance between the plates exceeds 2 cm. The
decolorization rate of 2.5 cm plate spacing increased the fastest. The decolorization of methyl orange first increased
and then decreased with the plate spacing. When the distance between the plates was 1 and 1.5 cm, the results of the
experiment within 10 min was relatively good. After that, the decolorization rate increased slowly. Considering that the
polar plate spacing was too small, and the liquid flow rate was too high, the floc body cannot grow steadily. When the
polar plate spacing was 2, 2.5, and 3 cm, the difference between the experimental results was not large. From Fig. 4, it
can be seen that the 3 cm spacing decolorization rate is lower than the previous two. Compared with 2 and 2.5 cm, the
result was not stable, and the decolorization rate was lower than 2.5 cm. Therefore, the more stable and better result of
2.5 cm interval was selected as the best plate spacing.
5
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Fig. 5. UV–visible measurements to assess the effect of plate angle on the experiment.

3.1.3. Effect of the plate angle


In this reactor, the size of the plate angle is not only related to the material loss and energy consumption but also
affects the flow rate of waste liquid between plates, thus affecting the flocculation effect, which has a great impact on the
removal efficiency. Therefore, it is very important to explore the angle of the electrode plate. The group of the experiment
was conducted under the initial conditions of the edge length of 3 cm, the flow rate of 60 L h−1 , and the polar plate
spacing of 2.5 cm. The effects of plate folds from 45 to 105◦ on EC were explored. As shown in Fig. 5, the decolorization
rate increases rapidly within 30 min, reaches more than 90% in the 30th minute, and hardly increases after 30 min. After
the experiment 50 min, the results showed that with the increase of the plate angle, the removal of methyl orange was
reduced after the first increase, and the gap was within 1.5%. As shown from Fig. 5, the decolorization rate is about 90%
in the 30th minute, but the 45◦ folding plate is 25% higher than the effective area of the 75◦ folding plate, indicating that
energy consumption and material loss increased by 25%, but the decolorization rate changed little. The results of three
folded plates 75◦ , 90◦ , and 105◦ were similar. The decolorization rate of 50 min and 75◦ folded plate was the highest, so
the experiment was carried out after 75◦ thereafter.
The above results showed that when the flow rate was 60 L h−1 , the polar plate spacing was 2.5 cm, and the polar
plate angle was 75◦ , the decolorization rate of wastewater was the highest, at 99.23%.

3.2. RSM experiment

3.2.1. RSM results and tests


The decolorization rate of the EC treatment methyl orange synthetic wastewater was the response value, and the
experiment was carried out using the optimal parameters above. The test results are listed in Table 3. Using Design-Expert
test design software, the data in the table were multi-linear linear regression fitting analysis, and the response surface
function of methyl orange decolorization rate was obtained. Finally, the regression model was a second-order response
surface, as shown in the following formula (5).
Y = 0.84 − 0.076A + 0.053B + 0.052C + 0.009357AB + 0.033AC − 0.01BC + 0.04A2 − 0.009447B2 + 0.031C 2 (5)
F-value and p-value were calculated by ANOVA to evaluate the significance of the model and its items. When the p-
value calculated is lower than 0.05, the test variables are significant. As shown in Table 4, independent variables, including
the initial MO concentration (A), current density (B), and experimental time (C), significantly affect the removal efficiency.
The interaction of AB and BC were insignificant for the experiment, while the interaction of AC was significant for the
experiment. The secondary effects of A and C had a significant influence on the reaction, while the secondary effect
of B had an insignificant influence on the experiment (P-value < 0.05,F-value ≥ 10.96). The Predicted R2 =0.8639 is in
reasonable agreement with the Adjusted R2 =0.9739; the difference is less than 0.2. Adeq Precision measures the signal to
noise ratio. A ratio greater than 4 is desirable. Adeq Precision=26.880 indicates an adequate signal. This model can be used
to navigate the design space. According to the parameters P-value < 0.0001 and R2 =0.9886 obtained from the experiment,
it can conclude that the prediction model is well developed and suitable for describing experimental data.
The normal plot of the residuals is depicted in Fig. 6. All residuals are well distributed along the straight line and show
the normal distribution of the obtained MO removal efficiency results. In addition, as another confirmation to support
data normality, Fig. 7 shows the predicted values and actual values of removal efficiency. Alternatively, since the values
obtained by R2 and Adj.R2 are close to each other, they support the normal state of the results obtained above.
6
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Table 3
Results of the experiment.
The experiment number Methyl orange concentration (mg L−1 ) Current density (mA cm−2 ) Reaction time (min) Decoloring rate
1 300 10 30 75.43%
2 100 5 40 89.34%
3 200 10 40 84.54%
4 100 10 50 99.23%
5 300 5 40 72.23%
6 200 10 40 83.54%
7 100 15 40 99.12%
8 200 10 40 82.45%
9 200 10 40 84.34%
10 300 15 40 85.76%
11 200 5 50 87.98%
12 200 5 30 73.87%
13 300 10 50 90.56%
14 200 15 50 95.45%
15 100 10 30 97.35%
16 200 10 40 82.74%
17 200 15 30 85.35%

Table 4
ANOVA results for response surface of quadratic model.
Source Sum of squares df Mean square F-value p-value
Model 0.106884 9 0.011876 67.45297 <0.0001 Significant
A—concentration. 0.046604 1 0.046604 264.6995 <0.0001
B—current. 0.022324 1 0.022324 126.7939 <0.0001
C—Time. 0.021239 1 0.021239 120.63 <0.0001
AB 0.000352 1 0.000352 1.996788 0.2005
AC 0.004389 1 0.004389 24.92879 0.0016
BC 0.000402 1 0.000402 2.283275 0.1745
A2 0.006856 1 0.006856 38.94097 0.0004
B2 0.000376 1 0.000376 2.134517 0.1874
C2 0.004008 1 0.004008 22.76389 0.002
Residual 0.001232 7 0.000176
Lack of Fit 0.000886 3 0.000295 3.407107 0.1336 Not significant
Total Color 0.108117 16

R2 = 0.9886 Adjusted R2 = 0.9739 Predicted R2 = 0.8639 Adeq Presion=26.8798.

Fig. 6. Normal probability plot of studentized residuals of the model for removal efficiency.

The Pareto analysis was used to evaluate the importance of the major variables and their interaction with the removal
efficiency, according to the following equation (Ahmadzadeh and Dolatabadi, 2018; Alidadi et al., 2018; Dolatabadi and
7
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Fig. 7. Experimental response values versus predicted response value.

Fig. 8. Pareto plot of main variables effects on removal efficiency.

Ahmadzadeh, 2019; Dolatabadi et al., 2019):


α2
Pi = ∑ i × 100(i ̸= 0)
αi2
Where Pi represents the percentage effect of each variable and αi represents the statistical significance coefficient in the
equation.
It can be seen from Fig. 8 that the initial concentration of MO and the reaction time are the most decisive factors that
synergistically influence the removal efficiency of MO as high as 55.8%. In addition, the contribution of B, AC, A2 and C2
to MO removal efficiency was 18.5%, 7.2%, 10.5% and 6.3%, respectively.

3.2.2. Analysis of experimental results


As can be known from Eq. (5) and Table 4, the decolorization rate of wastewater is not only influenced by a single factor
such as the initial concentration of methyl orange, current density, and reaction time. The effects of each factor on the
decolorization rate ranged from large to small: methyl orange concentration, current density, reaction time, the interaction
between methyl orange concentration, and reaction time. There was no interaction between current density and reaction
time, methyl orange concentration, and current density. However, it is necessary to study the optimal experimental
conditions for each factor.
8
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Fig. 9. The relationship between the decolorization rate of wastewater and the initial concentration and reaction time of methyl orange, the current
density is 10 ma cm−2 , and the other conditions are the best parameters in 3.4.

The effect of reaction time on removal efficiency at different methyl orange concentrations was studied, as shown in
Fig. 9. The current density of 10 mA cm −2 remained the same, and the change of the initial concentration of methyl orange
from 30–50 min at the reaction time was studied. It can be seen that at the methyl orange concentration of 200 mg L−1 , the
decolorization rate of methyl orange wastewater increased from 81.47% to 91.77% when the reaction time was increased
from 30 min to 50 min. The decolorization rate curve observed from the figure indicates that the optimal range of the
initial concentration of methyl orange is 100–250 mg L−1 , the optimal range of the initial concentration of methyl orange
within 40 min is about 100–140 mg L−1 and after 250 mg L−1 , the decolorization rate is inversely related to the previous
concentration due to the high concentration.
Fig. 10 shows the effect of current density and reaction time on the decolorization rate. Under the condition that the
initial concentration of methyl orange was 200 mg L−1 , the change of current density 5–15 mA cm−2 and reaction time
from 30–50 min of wastewater decolorization rate was studied. As shown in the figure, when factor B moves to the peak,
the contour density is significantly higher than that along with factor C, indicating that the current density exerts a greater
effect on the response value, which is consistent with the result of ANOVA. When the current density is lower than 11
mA cm−2 , the contour density is greater than that above 11 mA cm−2 , indicating that the effect of the current density on
the response value is greater in the case of the current density below 11 mA cm−2 , and the effect of the current density
on the response value is more significant when the reaction time is longer. At 40 min of reaction time, when the current
density increases from 5 to 15 mA cm−2 , the decolorization rate of methyl orange wastewater increases from 77.3% to
86.81%. The decolorization rate curve observed from the figure indicates that 30–35 min is the optimal reaction period
between 5–15 mA cm−2 and the optimum time is approximately 33 min.
Fig. 11 shows the effect of current density and methyl orange concentration at 40 min on the decolorization rate.
The change of the decolorization rate of methyl orange wastewater was studied under the conditions of the initial
concentration of 100–300 mg L−1 and the current density of 5–15 mA· cm−2 . The contour density of factor A is higher
than that of factor B, indicating that the initial concentration of MO has a greater impact on the response value, which is
consistent with the results of ANOVA. When the MO concentration is lower than 200 mg L−1 , the contour density is greater
than that of MO concentration above 200 mg L−1 , indicating that the initial MO concentration is lower than 200 mg L−1 ,
and the impact of the initial MO concentration on the response value is more significant when the current density is lower.
It can be found that at the initial concentration of methyl orange at 200 mg L−1 , the decolorization rate of methyl orange
wastewater increased from 77.03% to 87.88% by increasing the current density from 5 to 15 mA· cm−2 . As shown from the
figure, a direct relationship between the increase in current density and the decolorization rate of synthetic wastewater
is depicted. The decolorization rate curve indicates that the optimal range of the initial concentration of methyl orange is
approximately 100–250 mg L−1 , and the range of the initial concentration of methyl orange with a decolorization rate of
more than 90% is approximately 100–170 mg L−1 . As the current density increases, so does the decolorization rate, which
can result from the increasing number of Fe3+ plus ions produced in place, thereby controlling the growth of anodes and
cathodes in the EC process at the rate of reaction adsorption of contaminants. When the MO concentration is greater than
280 mg L−1 , the decolorization rate is inversely related to the previous concentration due to the high concentration.
The purpose of exploring the three elements A, B, and C is to explore the influence of the three factors on the
experiment and find out the optimal experimental conditions. According to the results of ANOVA, AB and BC have
9
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Fig. 10. The relationship between wastewater decolorization rate and current density and reaction time, the initial concentration of methyl orange
was 200 mg L−1 and the other conditions were the best parameters in 3.4.

Fig. 11. The relationship between the decolorization rate of wastewater and the current density and the initial concentration of methyl orange; the
reaction time is 40 min, and the other conditions are the best parameters in 3.4.

no significant influence on the response value, implying that AB and BC have no interaction. This means that MO
concentration (A) has little influence on the current density (B), and current density (B) has no effect on the reaction
time (C), in which MO concentration and density are independent, and density and reaction time are independent.

3.2.3. Optimal results and validation


To improve the work efficiency, the time was set to 30 min. Through the analysis of the Box–Behnken design model,
the maximum decolorization rate of the process parameters obtained at the maximum MO concentration and minimum
current density were: methyl orange concentration 134.03 mg L−1 , current density 10.09 mA cm−2 , and other conditions
remain unchanged. Under the optimized conditions, the decolorization rate of dyeing wastewater is estimated to be
90.53%. To verify the reliability of the model, according to the actual operation, the optimized process parameters were
modified as methyl orange concentration 134 mg L−1 , current density 10.1 mA cm−2 , reaction time 30 min, and other
conditions to optimize. In 3 sets of parallel experiments under this parameter, the decolorization rate of dyeing wastewater
was measured at 92.26%, 93.22%, and 91.57%, respectively, and the average decolorization rate was 92.35%, while flat
electrode under the same conditions of methyl orange decolorization rate was only 58.9%. The prediction value of the
10
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Table 5
Dynamic model parameters for removing methyl orange under optimized EC process conditions.
First-order dynamics R2 0.9887
K1 0.3315
Second-order dynamics R2 0.9954
K2 0.00189
qe 74.046

model is controlled at about 3%, and it is higher than the prediction value, which shows that the repetition of the
experiment is good. The model can better reflect the relationship between the decolorization rate of dyeing wastewater
and the selected factors.

3.3. Dynamics research

Adsorption dynamics are commonly used for the first and second-order dynamic models (Ahsan et al., 2018; Das
et al., 2017; Essandoh et al., 2021; Jamali-Behnam et al., 2018), representing different adsorption methods. The first-
order adsorption models are mainly diffusion adsorption, and the second-order adsorption models are mainly chemical
adsorption. The optimal conditions of the current operation, methyl orange concentration of 134 mg L−1, and current
density of 10.1 mA cm−2 were studied. The optimal parameters were used for experiments.
The pseudo-first-order kinetic model is shown in the formula.

log (qe − qt ) = logqe − k1 t /2.303 (6)


−1 −1
Where qe (mg g ) and qt (mg g ) are the balanced adsorption capacity of the t (min) moment. k1 is the first order
adsorption constant. The plot of log (qe - qt ) and t should be the linear relationship, therefrom k1 and qe can be determined
by the slope and intercept respectively. Under all concentrations studied, the results show that the theoretical qe (cal)
values agree with the experimental qe (exp) values. As can be seen from Fig. 12, the correlation coefficient between log
(qe - qt ) and t is 0.9887. The dynamic data is fitted to the first-order model equation and the calculated k1 is displayed
in Table 5. Next, the second-order Lagergren model is used to fit the experimental data. The pseudo-second-order kinetic
model is shown in the formula.
t 1 t
= + (7)
qt k2 q2e qe
Where qe (mg g−1 ) and qt (mg g−1 ) are the balanced adsorption capacity of the t (min) moment. k2 is the second-order
adsorption constant. Dynamic data fits into second-order model equations. Fig. 13 shows the curves of t/qt and t when
MO is adsorbed. The results show that these graphs have good linear correlation coefficients. Under all current densities
studied, the theoretical qe (cal) values are consistent with the experimental qe (exp) values. This indicates that the second-
order model agrees well with the experimental data and can explain the adsorption of MO on Fe(OH)3 . As shown in Fig. 13,
the correlation coefficient between t/qt and t is 0.9954 to calculate the equilibrium adsorption capacity qe and k2 and is
summarized in Table 5.
From the above data, the experimental results show that compared with the first-order model of methyl orange
adsorption dynamics on the upper Fe(OH)2 plus, the second-order model is more consistent with the experimental results.

3.4. Study of iso-temperature model

Metal hydroxides are produced as adsorbents during EC. The adsorption of pollutants on the surface of the resulting
metal hydroxide was investigated by two common adsorptions continues, Freundlich and Langmuir models. Through the
analysis of the adsorption iso-temperature line, the structure of the adsorption layer and the size of the adsorption force
can be obtained, and the adsorption mechanism can be further mastered.
Freundlich iso-temperature adsorption equations are one of the most widely used to describe adsorption balance.
It is mainly used to describe adsorption under non-ideal conditions and multi-molecular layer adsorption. Freundlich
iso-absorption equation as an experimental adsorption iso-temperature equation of uneven surface, which can not only
describe the adsorption process of an uneven surface but also make a reasonable explanation of the experimental results
in a wider concentration range.
The linear model of the Freundlich iso-temperature line is formulated as follows (Mahmoud et al., 2020; Singh et al.,
2018; Sirajudheen et al., 2020):
1
logqe = logkf + (8)
nlogce
where qe is the equilibrium adsorption, kf is the Freundlich constant, n is the constant associated with the adsorption
strength, and ce is the equilibrium concentration.
11
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Fig. 12. Effect of time on degradation of MO and fit of kinetics data with the linear plot of pseudo-first-order.

Fig. 13. Effect of time on degradation of MO and fit of kinetics data with the linear plot of pseudo-second-order.

Through Eq. (8), draw the relationship between the paired number qe and the logarithm ce of the adsorption data,
and the result should be a straight line with a slope of n and an intercept of kf (see Fig. 14). The intercept and slope
are the indexes of adsorption capacity and adsorption strength, respectively. N value in the range of 1 to 10 indicates
good adsorption. Table 6 lists the kf and n values for each concentration and current density. The results show that the
Freundlich diagram does not agree with the experimental data.
The Langmuir iso-temperature model is based on the Langmuir single molecular layer adsorption theory, which covers
the extent of molecules covering the solid surface of adsorbent to the extent of aggregation.
The linear model of the Langmuir iso-temperature line is represented as the following formula (Eltaweil et al., 2020;
Salimpour Abkenar et al., 2015; Tambat et al., 2020):
ce 1
= + ce qmax (9)
qe kl qmax
Where ce is the equilibrium concentration, qe is the equilibrium adsorption, kl is the Langmuir constant, and qmax is the
saturation adsorption. Generally, RL is adopted to describe the equilibrium constant or outline separation constant of the
Langmuir iso-temperature model, which can be described as follows:

Rl = 1/(1 + kl c0 ) (10)

Where c0 is the initial concentration, and kl is the Langmuir constant. The larger the qmax . indicates a stronger adsorption
capacity, and qmax is between 0 and 1. The Rl value indicates favorable adsorption. The experimental conditions used in
12
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Table 6
Langmuir and Freundlich adsorption contain models are used to remove MO EC values under optimized EC process
conditions.
Langmuir iso-warm line Freundlich iso-temperature line
qm (mg g−1 ) kl (L mg−1 ) RL R2 Kf N R2
83. 82 9.3937 7.94 × 10−4 0.99128 51.91 20.47 0.98054

Fig. 14. Equilibrium isotherm Freundlich of degradation of MO dye.

Fig. 15. Equilibrium isotherm Langmuir of degradation of MO dye.

this modeling are the best parameters for this study. As shown in Fig. 15, the correlation coefficient between curve ce /qe
and ce is 0.9912. The resulting slope represents the maximum adsorption capacity, respectively. The values of kl and qmax
are listed in Table 6.
Table 6 lists the relevant coefficient values for single-tier capacity qmax , Langmuir constant (kl ), equilibrium constant
(RL ), and two isometrical models. It enables us to find that the Langmuir iso-temperature model has a higher regression
coefficient (R2 :0.99128) than the Freundlich model, implying that the Langmuir model exerts a better process description.
The results show that MO is single-layer adsorption on the adsorbent surface.

3.5. Process comparison

Several processes for treating dyes are listed in Table 7. It can be seen that these processes have high degradation
rates for dyes, but they also have obvious disadvantages, such as the long degradation cycle of biological methods.
13
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Table 7
Several processes for treating dyes.
Treatment process Dye type Concentration (mg L−1 ) Processing time (min) Decolorization rate References
Microbiological method Orange II 300 360 78.2% Suzuki et al. (2020)
Electrocatalysis RRX-3B 1650 90 100% Lu et al. (2021)
photocatalysis MG 30 45 81% Iqbal et al. (2021)
adsorption MB 5 20 99.5% Kubra et al. (2021)
adsorption MB 400 60 99% Li et al. (2021)

Electrocatalytic need to add hydrogen peroxide, the cost is relatively high. Photocatalytic degradation is not complete.
The adsorption process requires the preparation of new adsorbents, the process is complex and the cost is high. The
electrocoagulation process is simple, and the cost is low. It is the ultimate goal to develop a continuous high-performance
electrocoagulation reactor. This experiment is the initial study.

3.6. Economic evaluation

Cost analysis is vital in wastewater treatment technology, which restricts the application of industry to a great extent.
The costs related to EC contain three parts: the cost of energy consumption, the cost of dissolving electrodes, and the cost
of adding any chemicals (to increase the conductivity of the solution or to change the pH of the solution). The operating
cost of using EC can be depicted mathematically.

3.6.1. Electrode consumption


Formulate the electrode consumption (ELC) mathematically as follows:
ItM
ELC = (11)
nF
Where F, M, t, n, and I express Faraday constants (96,485 C mol−1 ), molecular weight (g mol−1 ), reaction time (s), electrons,
and applied current (A). Taking into account the best conditions, the ELC for a single run is 0.2376g. Flat electrode.

3.6.2. Energy consumption


Calculate the energy consumption in the experiment by using the formula:
UIt
EEC =
V
Where V, t, I, and U represent sample volume (L), reaction time (h), applied current (A), and applied voltage (V). Taking
into account the optimum conditions, the measured energy consumption is 1.5925 (kWh m−3 ).

4. Conclusion

In this study, a continuous electrocoagulation reactor was designed to treat methyl orange synthesis wastewater with
a folding plate electrode. At a pH of 6, by single factor experiment analysis, the optimal conditions were flow rate of 60
L h−1 , electrode spacing of 2.5 cm, and folding Angle of 75◦ . The decolorization rate was 99.23% in 50 min. The response
surface methodology experiment showed that the initial concentration of methyl orange had the greatest influence on
the experiment. The optimal conditions were as follows: the initial concentration of methyl orange was 134 mg L−1 , the
current density was 10.1 mA cm−2 , and the removal rate was 92.35% in 30 min. For the removal of methyl orange synthetic
wastewater, the results showed that the electrode had a loss of 0.2376g per experiment, and the power consumption
required for a single operation is 1.5925 kWh m−3 . Dynamics studies have shown that the adsorption of methyl orange on
iron hydroxide can be described using second-order dynamics models. Methyl orange was more suitable for the Langmuir
adsorption iso-temperature line, and the results were good in line with the experimental data.
In this study, the liquid flow rate, plate spacing, and plate Angle in the reactor were discussed, and the plate side length
was not discussed. In future research, the length of the plate side, the number of the plate, and the way of multiple reactors
in series will be studied, and a set of continuous treatment electrocoagulation reactor will be developed.

CRediT authorship contribution statement

Ze Wu: Methodology, Data curation, Writing - original draft. Jie Dong: Software. Yingwu Yao: Supervision. Yang
Yang: Supervision, Validation. Feng Wei: Conceptualization, Visualization, Supervision Validation, Writing - reviewing
and editing.
14
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

References

Abbasi, S., Mirghorayshi, M., Zinadini, S., Zinatizadeh, A.A., 2020. A novel single continuous electrocoagulation process for treatment of licorice
processing wastewater: Optimization of operating factors using RSM. Process Saf. Environ. Prot. 134, 323–332. http://dx.doi.org/10.1016/j.psep.
2019.12.005.
Ahmadzadeh, S., Dolatabadi, M., 2018. In situ generation of hydroxyl radical for efficient degradation of 2, 4-dichlorophenol from aqueous solutions.
Environ. Monit. Assess. 190, 340. http://dx.doi.org/10.1007/s10661-018-6697-0.
Ahmadzadeh, S., Kassim, A., Rezayi, M., et al., 2011. A Conductometric Study of Complexation Reaction Between Meso-octamethylcalix[4]pyrrole with
Titanium Cation in Acetonitrile–Ethanol Binary Mixtures. Int. J. Electrochem. Sci. 6 (10), http://dx.doi.org/10.1002/fuce.201000173.
Ahmadzadeh, S., Rezayi, M., Faghih-Mirzaei, E., Yoosefian, M., Kassim, A., 2015. Highly selective detection of titanium (III) in industrial waste
water samples using meso-octamethylcalix[4]pyrrole-doped PVC membrane ion-selective electrode. Electrochim. Acta 178, 580–589. http:
//dx.doi.org/10.1016/j.electacta.2015.07.014.
Ahmadzadeh, S., Yoosefian, M., Rezayi, M., 2019. Comprehensive experimental and theoretical investigations on chromium (III) trace detection in
biological and environmental samples using polymeric membrane sensor. Int. J. Environ. Anal. Chem. 1–16. http://dx.doi.org/10.1080/03067319.
2019.1685664.
Ahsan, M.A., et al., 2018. Adsorptive removal of methylene blue, tetracycline and Cr(VI) from water using sulfonated tea waste. Environ. Technol.
Innov. 11, 23–40. http://dx.doi.org/10.1016/j.eti.2018.04.003.
Al-Raad, A.A., Hanafiah, M.M., Naje, A.S., Ajeel, M.A., 2020. Optimized parameters of the electrocoagulation process using a novel reactor with rotating
anode for saline water treatment. Environ. Pollut. 265, 115049. http://dx.doi.org/10.1016/j.envpol.2020.115049.
Alidadi, H., et al., 2018. Enhanced removal of tetracycline using modified sawdust: Optimization, isotherm, kinetics, and regeneration studies. Process
Saf. Environ. Prot. 117, 51–60. http://dx.doi.org/10.1016/j.psep.2018.04.007.
Ammar, S.H., Ismail, N.N., Ali, A.D., Abbas, W.M., 2019. Electrocoagulation technique for refinery wastewater treatment in an internal loop split-plate
airlift reactor. J. Environ. Chem. Eng. 7, http://dx.doi.org/10.1016/j.jece.2019.103489.
Avazpour, S., Pardakhty, A., Nabatian, E., Ahmadzadeh, S., 2020. Economical approach for determination of kojic acid by nanostructured ionic
liquid-based carbon paste sensor. BioNanoScience 10, 502–511. http://dx.doi.org/10.1007/s12668-020-00723-3.
Ba Mohammed, B., et al., 2020. Fe-ZSM-5 zeolite for efficient removal of basic Fuchsin dye from aqueous solutions: Synthesis, characterization and
adsorption process optimization using BBD-RSM modeling. J. Environ. Chem. Eng. 8, http://dx.doi.org/10.1016/j.jece.2020.104419.
Badakhshan, S., Ahmadzadeh, S., Mohseni-Bandpei, A., Aghasi, M., Basiri, A., 2019. Potentiometric sensor for iron (III) quantitative determination:
experimental and computational approaches. BMC Chem. 13, 131. http://dx.doi.org/10.1186/s13065-019-0648-x.
Barhoumi, A., et al., 2019. High-rate humic acid removal from cellulose and paper industry wastewater by combining electrocoagulation process
with adsorption onto granular activated carbon. Ind. Crop. Prod. 140, http://dx.doi.org/10.1016/j.indcrop.2019.111715.
Bener, S., et al., 2019. Electrocoagulation process for the treatment of real textile wastewater: Effect of operative conditions on the organic carbon
removal and kinetic study. Process. Saf. Environ. Prot. 129, 47–54. http://dx.doi.org/10.1016/j.psep.2019.06.010.
Brillas, E., Martínez-Huitle, C.A., 2015. Decontamination of wastewaters containing synthetic organic dyes by electrochemical methods. An updated
review. Appl. Catal. B 166–167, 603–643. http://dx.doi.org/10.1016/j.apcatb.2014.11.016.
Combatt, M.P.M., et al., 2020. Design of parallel plate electrocoagulation reactors supplied by photovoltaic system applied to water treatment. Comput.
Electron. Agric. 177, http://dx.doi.org/10.1016/j.compag.2020.105676.
Das, R., Bhaumik, M., Giri, S., Maity, A., 2017. Sonocatalytic rapid degradation of congo red dye from aqueous solution using magnetic Fe(0)/polyaniline
nanofibers. Ultrason. Sonochem. 37, 600–613. http://dx.doi.org/10.1016/j.ultsonch.2017.02.022.
Dolatabadi, M., Ahmadzadeh, S., 2019. A rapid and efficient removal approach for degradation of metformin in pharmaceutical wastewater using
electro-fenton process; optimization by response surface methodology. Water Sci. Technol. 80, 685–694. http://dx.doi.org/10.2166/wst.2019.312.
Dolatabadi, M., Ahmadzadeh, S., Ghaneian, M.T., 2019. Mineralization of mefenamic acid from hospital wastewater using electro-fenton degradation:
Optimization and identification of removal mechanism issues. Environ. Prog. Sustain. Energy 39, http://dx.doi.org/10.1002/ep.13380.
Dolatabadi, M., Mehrabpour, M., Esfandyari, M., Ahmadzadeh, S., 2020. Adsorption of tetracycline antibiotic onto modified zeolite: Experimental
investigation and modeling. MethodsX 7, http://dx.doi.org/10.1016/j.mex.2020.100885.
Eltaweil, A.S., Ali Mohamed, H., Abd El-Monaem, E.M., El-Subruiti, G.M., 2020. Mesoporous magnetic biochar composite for enhanced adsorption
of malachite green dye: Characterization adsorption kinetics, thermodynamics and isotherms. Adv. Powder Technol. 31, 1253–1263. http:
//dx.doi.org/10.1016/j.apt.2020.01.005.
Essandoh, M., Garcia, R.A., Palochik, V.L., Gayle, M.R., Liang, C., 2021. Simultaneous adsorption of acidic and basic dyes onto magnetized
polypeptidylated-hb composites. Sep. Purif. Technol. 255, http://dx.doi.org/10.1016/j.seppur.2020.117701.
Fan, T., Deng, W., Feng, X., Pan, F., Li, Y., 2020. An integrated electrocoagulation - electrocatalysis water treatment process using stainless steel
cathodes coated with ultrathin tio2 nanofilms. Chemosphere 254, 126776. http://dx.doi.org/10.1016/j.chemosphere.2020.126776.
Feng, C., et al., 2020. Facile synthesis of trimethylammonium grafted cellulose foams with high capacity for selective adsorption of anionic dyes from
water. Carbohydr. Polym. 241, 116369. http://dx.doi.org/10.1016/j.carbpol.2020.116369.
Hogeboom, R.J., 2020. The water footprint concept and water’s grand environmental challenges. One Earth 2, 218–222. http://dx.doi.org/10.1016/j.
oneear.2020.02.010.
Iqbal, M., et al., 2021. Graphene oxide nanocomposite with cuse and photocatalytic removal of methyl green dye under visible light irradiation.
Diam. Relat. Mater. 113, http://dx.doi.org/10.1016/j.diamond.2021.108254.
Jamali-Behnam, F., et al., 2018. Adsorptive removal of arsenic from aqueous solutions using magnetite nanoparticles and silica-coated magnetite
nanoparticles. Environ. Prog. Sustain. Energy 37, 951–960. http://dx.doi.org/10.1002/ep.12751.
Khandegar, V., Saroha, A.K., 2013. Electrocoagulation for the treatment of textile industry effluent–a review. J. Environ. Manage. 128, 949–963.
http://dx.doi.org/10.1016/j.jenvman.2013.06.043.
Khorram, A.G., Fallah, N., 2018. Treatment of textile dyeing factory wastewater by electrocoagulation with low sludge settling time: Optimization of
operating parameters by RSM. J. Environ. Chem. Eng. 6, 635–642. http://dx.doi.org/10.1016/j.jece.2017.12.054.
Kim, T., Kim, T.-K., Zoh, K.-D., 2020. Removal mechanism of heavy metal (Cu, Ni, Zn, and Cr) in the presence of cyanide during electrocoagulation
using Fe and Al electrodes. J. Water Process. Eng. 33, http://dx.doi.org/10.1016/j.jwpe.2019.101109.
Kong, Z., Li, L., Xue, Y., Yang, M., Li, Y.-Y., 2019. Challenges and prospects for the anaerobic treatment of chemical-industrial organic wastewater: A
review. J. Cleaner Prod. 231, 913–927. http://dx.doi.org/10.1016/j.jclepro.2019.05.233.

15
Z. Wu, J. Dong, Y. Yao et al. Environmental Technology & Innovation 22 (2021) 101448

Kubra, K.T., Salman, M.S., Hasan, M.N., 2021. Enhanced toxic dye removal from wastewater using biodegradable polymeric natural adsorbent. J.
Molecular Liquids 328, http://dx.doi.org/10.1016/j.molliq.2021.115468.
Kuppusamy, S., et al., 2017. Quercus robur acorn peel as a novel coagulating adsorbent for cationic dye removal from aquatic ecosystems. Ecol. Eng.
101, 3–8. http://dx.doi.org/10.1016/j.ecoleng.2017.01.014.
Li, L., et al., 2021. Efficient removal of cationic dyes via activated carbon with ultrahigh specific surface derived from vinasse wastes. Bioresour.
Technol. 322, 124540. http://dx.doi.org/10.1016/j.biortech.2020.124540.
Lu, J., Ayele, B.A., Liu, X., Chen, Q., 2021. Electrochemical removal of RRX-3B in residual dyeing liquid with typical engineered carbonaceous cathodes.
J. Environ. Manage. 280, 111669. http://dx.doi.org/10.1016/j.jenvman.2020.111669.
Mahmoud, M.E., Abdelfattah, A.M., Tharwat, R.M., Nabil, G.M., 2020. Adsorption of negatively charged food tartrazine and sunset yellow dyes onto
positively charged triethylenetetramine biochar: Optimization, kinetics and thermodynamic study. J. Molecular Liquids 318, http://dx.doi.org/10.
1016/j.molliq.2020.114297.
Malakootian, M., Moridi, A., 2017. Efficiency of electro-fenton process in removing acid red 18 dye from aqueous solutions. Process. Saf. Environ.
Prot. 111, 138–147. http://dx.doi.org/10.1016/j.psep.2017.06.008.
Mirbolooki, H., Amirnezhad, R., Pendashteh, A.R., 2017. Treatment of high saline textile wastewater by activated sludge microorganisms. J. Appl. Res.
Technol. 15, 167–172. http://dx.doi.org/10.1016/j.jart.2017.01.012.
Moussa, D.T., El-Naas, M.H., Nasser, M., Al-Marri, M.J., 2017. A comprehensive review of electrocoagulation for water treatment: Potentials and
challenges. J. Environ. Manag. 186, 24–41. http://dx.doi.org/10.1016/j.jenvman.2016.10.032.
Nasrullah, M., Singh, L., Krishnan, S., Sakinah, M., Zularisam, A.W., 2018. Electrode design for electrochemical cell to treat palm oil mill effluent by
electrocoagulation process. Environ. Technol. Innov. 9, 323–341. http://dx.doi.org/10.1016/j.eti.2017.10.001.
Nguyen, C.H., Juang, R.-S., 2019. Efficient removal of cationic dyes from water by a combined adsorption-photocatalysis process using platinum-doped
titanate nanomaterials. J. Taiwan Inst. Chem. Eng. 99, 166–179. http://dx.doi.org/10.1016/j.jtice.2019.03.017.
Nunez, J., et al., 2019. Application of electrocoagulation for the efficient pollutants removal to reuse the treated wastewater in the dyeing process
of the textile industry. J. Hazard. Mater. 371, 705–711. http://dx.doi.org/10.1016/j.jhazmat.2019.03.030.
Ozbey Unal, B., et al., 2019. Adsorption and fenton oxidation of azo dyes by magnetite nanoparticles deposited on a glass substrate. J. Water Process.
Eng. 32, http://dx.doi.org/10.1016/j.jwpe.2019.100897.
Ozyonar, F., Gokkus, O., Sabuni, M., 2020. Removal of disperse and reactive dyes from aqueous solutions using ultrasound-assisted electrocoagulation.
Chemosphere 258, 127325. http://dx.doi.org/10.1016/j.chemosphere.2020.127325.
Rawat, D., Mishra, V., Sharma, R.S., 2016. Detoxification of azo dyes in the context of environmental processes. Chemosphere 155, 591–605.
http://dx.doi.org/10.1016/j.chemosphere.2016.04.068.
Rezayi, M., Ahmadzadeh, S., Kassim, A., et al., 2011a. Thermodynamic Studies of Complex Formation Between Co(SALEN) Ionophore with Chromate
(II) Ions in AN-H 2 O Binary Solutions by The Conductometric Method. Int. J. Electrochem. Sci. 6 (12), 6350–6359.
Rezayi, M., et al., 2011b. Conductometric determination of formation constants of tris(2-pyridyl) methylamine and titanium (III) in water-acetonitryl
mixture. Int. J. Electrochem. Sci. 6, 4378–4387.
Rodrigues, A.R., Seki, C.C., Ramalho, L.S., Argondizo, A., Silva, A.P., 2020. Electrocoagulation in a fixed bed reactor – color removal in batch and
continuous mode. Sep. Purif. Technol. 253, http://dx.doi.org/10.1016/j.seppur.2020.117481.
Rounaghi, G.H., Mohajeri, M., Ahmadzadeh, S., Tarahomi, S., 2009. A thermodynamic study of interaction of Na+ cation with benzo-15-crown-5 in
binary mixed non-aqueous solvents. J. Incl. Phenom. Macrocycl. Chem. 63, 365–372. http://dx.doi.org/10.1007/s10847-009-9530-0.
Salimpour Abkenar, S., Malek, R.M.A., Mazaheri, F., 2015. Dye adsorption of cotton fabric grafted with PPI dendrimers: Isotherm and kinetic studies.
J. Environ. Manag. 163, 53–61. http://dx.doi.org/10.1016/j.jenvman.2015.08.003.
Samsami, S., Mohamadi, M., Sarrafzadeh, M.-H., Rene, E.R., Firoozbahr, M., 2020. Recent advances in the treatment of dye-containing wastewater
from textile industries: Overview and perspectives. Process. Saf. Environ. Prot. 143, 138–163. http://dx.doi.org/10.1016/j.psep.2020.05.034.
Singh, N.B., Nagpal, G., Agrawal, S., Rachna, ., 2018. Water purification by using adsorbents: A review. Environ. Technol. Innov. 11, 187–240.
http://dx.doi.org/10.1016/j.eti.2018.05.006.
Sirajudheen, P., Karthikeyan, P., Vigneshwaran, S., Meenakshi, S., 2020. Synthesis and characterization of la(III) supported carboxymethylcellulose-
clay composite for toxic dyes removal: Evaluation of adsorption kinetics, isotherms and thermodynamics. Int. J. Biol. Macromol. 161, 1117–1126.
http://dx.doi.org/10.1016/j.ijbiomac.2020.06.103.
de Sousa Rollemberg, S.L., Mendes Barros, A.R., Milen Firmino, P.I., Bezerra Dos Santos, A., 2018. Aerobic granular sludge: Cultivation parameters and
removal mechanisms. Bioresour. Technol. 270, 678–688. http://dx.doi.org/10.1016/j.biortech.2018.08.130.
Su, C.X.-H., Low, L.W., Teng, T.T., Wong, Y.S., 2016. Combination and hybridisation of treatments in dye wastewater treatment: A review. J. Environ.
Chem. Eng. 4, 3618–3631. http://dx.doi.org/10.1016/j.jece.2016.07.026.
Suhadolnik, L., et al., 2019. Continuous photocatalytic, electrocatalytic and photo-electrocatalytic degradation of a reactive textile dye for wastewater-
treatment processes: Batch, microreactor and scaled-up operation. J. Ind. Eng. Chem. 72, 178–188. http://dx.doi.org/10.1016/j.jiec.2018.12.
017.
Suzuki, M., Suzuki, Y., Uzuka, K., Kawase, Y., 2020. Biological treatment of non-biodegradable azo-dye enhanced by zero-valent iron (ZVI)
pre-treatment. Chemosphere 259, 127470. http://dx.doi.org/10.1016/j.chemosphere.2020.127470.
Tambat, S.N., Ahirrao, D.J., Pandit, A.B., Jha, N., Sontakke, S.M., 2020. Hydrothermally synthesized N2-uio-66 for enhanced and selective adsorption
of cationic dyes. Environ. Technol. Innov. 19, http://dx.doi.org/10.1016/j.eti.2020.101021.
Tkaczyk, A., Mitrowska, K., Posyniak, A., 2020. Synthetic organic dyes as contaminants of the aquatic environment and their implications for
ecosystems: A review. Sci. Total Environ. 717, 137222. http://dx.doi.org/10.1016/j.scitotenv.2020.137222.
Villalobos-Lara, A.D., et al., 2020. CFD Simulation of biphasic flow, mass transport and current distribution in a continuous rotating cylinder electrode
reactor for electrocoagulation process. J. Electroanal. Soc. 858, http://dx.doi.org/10.1016/j.jelechem.2019.113807.
Wei, M.-C., et al., 2012. Improvement of textile dye removal by electrocoagulation with low-cost steel wool cathode reactor. Chem. Eng. J. 192,
37–44. http://dx.doi.org/10.1016/j.cej.2012.03.086.
Xu, L.L., et al., 2021. Development of a novel electrocoagulation membrane reactor with electrically conductive membranes as cathode to mitigate
membrane fouling. J. Membr. Sci. 618, http://dx.doi.org/10.1016/j.memsci.2020.118713.
Zaied, B.K., et al., 2020. A comprehensive review on contaminants removal from pharmaceutical wastewater by electrocoagulation process. Sci. Total
Environ. 726, 138095. http://dx.doi.org/10.1016/j.scitotenv.2020.138095.
Zhang, X., et al., 2015. Influence of sewage treatment on China’s energy consumption and economy and its performances. Renew. Sustain. Energy
Rev. 49, 1009–1018. http://dx.doi.org/10.1016/j.rser.2015.04.182.
Zhou, Y., Lu, J., Zhou, Y., Liu, Y., 2019. Recent advances for dyes removal using novel adsorbents: A review. Environ. Pollut. 252, 352–365.
http://dx.doi.org/10.1016/j.envpol.2019.05.072.

16

You might also like