Next Article in Journal
Integrating Socioeconomic Biophysical and Institutional Factors for Evaluating Small-Scale Irrigation Schemes in Northern Ethiopia
Previous Article in Journal
Bioactivities and Chemical Compositions of Cinnamomum burmannii Bark Extracts (Lauraceae)
Previous Article in Special Issue
Data Driven Modelling and Control Strategies to Improve Biogas Quality and Production from High Solids Anaerobic Digestion: A Mini Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Efficacy of Electrocoagulation Treatment for the Abatement of Heavy Metals: An Overview of Critical Processing Factors, Kinetic Models and Cost Analysis

by
Saif Ullah Khan
1,*,
Mohammad Khalid
1,2,*,
Khalid Hashim
3,
Mehdi Hassanvand Jamadi
4,5,
Milad Mousazadeh
6,7,
Farrukh Basheer
1 and
Izharul Haq Farooqi
1
1
Department of Civil Engineering, Zakir Husain College of Engineering & Technology, Aligarh Muslim University, Aligarh 202002, India
2
Department of Civil and Engineering, William States Lee College of Engineering, University of North Carolina Charlotte, Charlotte, NC 28223, USA
3
Built Environment and Sustainable Technologies Research Institute (BEST), Liverpool John Moores University, Byrom Street, Liverpool L3 3AF, UK
4
Omrab Canada Inc., 2283 Saint Laurent Blvd., Ottawa, ON K1G 5A2, Canada
5
Faculty of Civil, Water and Environmental Engineering, Shahid Beheshti University, Tehran 1983969411, Iran
6
Social Determinants of Health Research Center, Research Institute for Prevention of Non-Communicable Diseases, Qazvin University of Medical Sciences, Qazvin 3419915315, Iran
7
Department of Environmental Health Engineering, School of Health, Qazvin University of Medical Sciences, Qazvin 3419915315, Iran
*
Authors to whom correspondence should be addressed.
Sustainability 2023, 15(2), 1708; https://doi.org/10.3390/su15021708
Submission received: 31 October 2022 / Revised: 8 January 2023 / Accepted: 11 January 2023 / Published: 16 January 2023
(This article belongs to the Special Issue Solid-Waste and Waste-Water Treatment Processes)

Abstract

:
The electrocoagulation (EC) process introduces coagulants by electrochemical means, and is widely adopted for removing heavy metals, besides other contaminants, such as organic pollutants, suspended and dissolved solids, colloidal materials, etc. However, its capability can vary significantly, depending on the operating conditions. Although most of the investigations so far are limited at the laboratory level with artificially prepared solutions or industrial effluent lacking full- and field-scale studies, the success of the process depends a lot on optimizing the process variable. It has been found that the current density (typically 1–20 mA/cm2), type of electrode (generally aluminum or iron) and minimum electrolysis time are the key process parameters that influence performance. Furthermore, key mechanisms involved in the EC process, including charge neutralization, reduction-oxidation and precipitation/co-precipitation, are crucial for pollutant abatement. This review presents a detailed study undertaking all significant parameters that play a crucial role in the EC process, its mechanism, and improving the efficiency of this process by optimization of these parameters, along with suitable kinetic models.

1. Introduction

A clean and reasonable water supply to meet the needs of the growing global population has become a great challenge of this century [1,2]. Surface and groundwater, which are a major source of drinkable water supplies, besides desalinated water in water-scarce areas, are commonly contaminated by various pollutants, primarily heavy metals [3,4]. Increasing urbanization and expansion of industries, such as metallurgical industries, electroplating units, mining operations, electronic manufacturing units, fertilizer industries and leather industries, have resulted in discharging a considerable amount of wastewater containing high concentrations of trace metals, besides other contaminants, into the environment [5,6]. Several reports of heavy metal contamination of drinkable water supplies worldwide are available in the literature, of which a few are tabulated below in Table 1. There are numerous reports on health effects, due to the presence of these trace metals in water supplies [7,8]. Intake of chromium, lead, copper, nickel and cobalt may cause leukaemia, anaemia, hypertension, heart problems and cancer [9,10].
These toxic heavy metals cannot be easily removed from the wastewater without the use of advanced treatment technologies [28]. The practiced methods for removing heavy metals are adsorption, chemical precipitation and aquatic plants [29]. However, most of these have serious drawbacks, such as low efficiency, long treatment time or high operating cost. For example, advancements in the nano-technological field for water purification and treatment have also engrossed attention [30], although there are a few limitations, including high cost. Membrane filtration techniques have been widely used, but these also have disadvantages, such as membrane fouling and high cost. Therefore, effective and low-cost techniques need to be employed before discharging industrial effluents. Among conventional methods adopted, electrochemical-based approaches are currently used worldwide to remove metal ions because of their unique merits, such as the low consumption of chemicals, cost-effectiveness and simplicity. Electrocoagulation (EC), which is one of the electrochemical methods, has been successfully adopted for the remediation of wastewater from a wide range of pollutants, such as heavy metals [26], organic matter [31], textile dyes [32] and pharmaceutical wastewaters [33]. Preferring the EC method over conventional methods needs proper evaluation and in-depth analysis of various indices like a comparative assessment of the process based on operation compatibility, energy requirement, operational cost, sludge handling, etc. [34]. Therefore, there is a need for a feasible and cost-effective technology for the target removal of trace metals [35].
The EC method has certain advantages over other conventional methods. For example, although the chemical coagulation and EC methods have the same removal mechanism [35], the EC method is better than the chemical coagulation in terms of chemicals consumption, dissipation of coagulants [36], separation of flocs and sludge and sludge quality [37,38].
A few studies on heavy metal removal by conventional treatment methods on the comparative ground are shown in Table 2.
Electrochemical method-based water treatment was first noticed and reported long ago in 1889 in England. However, the first application of electrolysis was patented in 1904. Since then, treatment using EC has grabbed attention worldwide. On a larger scale, EC was first used in the USA, way back in 1946, for treating contaminants in drinking water. However, in the later half of the twentieth century, extensive work was undertaken to investigate the process mechanism involved. In the last two decades, the EC process has become quite popular, and it has been used in treating several types of wastewater and water, such as the removal of arsenic-containing smelting wastewater [45], mining and metal-processed wastewater [46], distillery-processed wastewater [47], nitrate-containing wastewater [48], paper and pulp industry wastewater [48], textile mill wastewater [49], tannery wastewater [50] and laundry wastewater [51]. Accordingly, numerous publications have stated the wide applicability of EC in removing heavy metal ions from water and wastewater [52]. These heavy metals include chromium, nickel, arsenic, cadmium, mercury, lead, copper, zinc, iron, etc. A sample of studies on optimizing and removing these heavy metal ions using the EC method has been tabulated in Table 3 below.
Although the EC method is a self-sufficient process in treating heavy metals with only a few exceptions, it is used as a hybrid process, along with some other treatment methods in a continuous phase [52]. However, several studies have used the EC method either as a pre-treatment or a polishing step. For instance, color and chemical oxygen demand (COD) removal from UASB (up-flow anaerobic sludge blanket) pre-treated poultry manure wastewater was carried out using EC as a polishing step. Under the highest substrate loading with HRT (Hydraulic Retention Time) of 8 days, the maximum COD removal attained was 81%, suggesting inadequate treatment by UASB alone as per environmental regulations. Similarly, a recent study conducted by Stylianou, et al. [25] showed significant removal of heavy metals. The results showed a 99.9% removal of Fe, Zn, Mn, and Cu, and 98%, 96%, and 88% for Ni, Cd and Cr, respectively. In another study conducted by Xu, et al. [27], keeping the initial pH of 6, they found a removal efficiency of 95.16% of copper. Thus, EC may be used as a polishing step for various types of wastewater, indicating its suitability to be applied as a preceding or polishing step.
The present work aims to review and analyze past research on heavy metal removal using the EC method. Additionally, it aims to define the role of different key parameters in EC treatment.

2. EC Theory and Mechanism

Overall, the EC method has a quite complex mechanism that includes various simultaneous physio-chemical reactions to generate the coagulants, as shown in Figure 1. The EC process involves in situ production of coagulating ions, with three consecutive phases: (i) production of coagulant by electrolytic oxidation of dissipating electrode (generally Al or Fe), (ii) destabilization of target pollutant, particulate suspension and (iii) aggregation of destabilized particles together to form flocs. The general mechanism undergoing inside an EC cell with dissipative electrodes of metal M can be described in the form of chemical reactions as follows:
At the anode:
M(s) → M(aq)n+ + ne
2H2O(l) → 4H+(aq) + O2(g) + 4e
At the cathode:
M(aq)n+ + ne → M(s)
2H2O(l) + 2e → H2(g) + 2OH
For iron electrodes, Fe(aq)3+ ions are formed by electrochemical oxidation, which further reacts to form resultant hydroxide [Fe(OH)3] and polyhydroxide species, namely: Fe(H2O)63+, Fe(H2O)5(OH)2+, Fe(H2O)4(OH)2+, Fe2(H2O)8(OH)24+ and Fe2(H2O)6(OH)42+, based on the pH conditions. These hydroxides/polyhydroxy metallic complexes show strong empathy towards distributed ions and particles, so they start coagulation. The gases evolved, resulting in the up-flowing of the coagulated particles. The reactions may be shown as follows:
At anode:
4Fe(s) → 4Fe2+(aq) + 8e
Precipitation:
4Fe2+ + 10H2O + O2 → 4Fe(OH)3 + 8H+
At cathode:
8H+ (aq) + 8e (aq) → 4H2(g)
Overall reaction:
4Fe(s) + 10H2O(l) + O2(g) → 4Fe(OH)3(s) ↓ + 4H2(g)
In addition to Fe3+ production, several reactions occur simultaneously in an EC system. These include:
  • formation of H2 gas at the cathode (Equation (4))
  • rise of pH as a result of hydroxyl ion formation (Equation (4))
  • reduction of metal ions on the cathodes.
Many researchers have proved the production of O2 at the anode according to Equation (2) above, though, this does not occur in usual conditions (electrochemical potential) in EC as anodic dissolution obeys Faraday’s rule [65]. However, in alkaline conditions, the dissolution of the iron anode is lesser than intended, as per Faraday’s law, which specifies that several other electrochemical reactions occur in these environments.

3. Factors Affecting EC and Their Significance

Most of the studies undertaken in the past have undertaken only a few parameters to optimize the EC process. However, to attain maximum removal efficiency with minimum energy consumption, various factors that influence the process, such as the initial pH, applied current, operational time, initial metal ion concentration, electrode material, electrode size, spacing and arrangement, the conductivity of the solution, agitation rate and nature of power supply (DC or AC). All these parameters have their particular role to play and need to be adjusted accordingly for the target removal of contaminants. An insight into these important factors and their role in EC is discussed in detail below.

3.1. Initial pH

pH plays a deterministic role in electrochemical coagulation. It is observed that pH does not remain still during the treatment, but mostly increases as reaction time proceeds [66]. The effect of pH depends on the type of pollutant, the material of anode used, and the reaction mechanism involved. Therefore, understanding the role of the pH in the reaction mechanism is important to achieve the best performance of the EC units [52]. Several studies have explored the role of pH in the performance of EC units. For instance, the effect of pH (from 2 to 10) on the removal of mercury, nickel and lead from solutions was investigated by [67]. The latter used a solution with an initial concentration of 10 mg/L of the mentioned metals, a current density of 0.15 mA/cm2, a treatment time of 30min, a temperature of 305 K and an anode and cathode made of magnesium and galvanized iron, respectively. Increasing the pH beyond the optimum value (7) negatively influenced the removal of Hg, Ni and Pb. A similar trend was noticed in the study of Kim, et al. [68], who used two different electrodes (aluminum and iron) to remove Cu, Ni, Zn and Cr from wastewater. The results showed that both electrodes sowed the best removal of the mentioned heavy metals at a neutral or slightly alkaline pH value. However, Xie, et al. [69] reported that the removal of silver using a graphite anode and aluminum cathode is not much influenced by the pH value, where it was noticed that the best removal of the silver was achieved after 20 min, regardless of the pH value. The outcomes of these studies confirmed the fact that the effect of pH on the performance of the EC units depends on several factors, including the type of electrodes and pollutants. Generally, the effects of the pH on the performance of the EC method are attributed to its impact on the species of the coagulants. The literature indicated that changes in the pH value led to the generation of different species of coagulants. For example, in iron electrodes, Fe2+ is the dominant specie when the pH < 8, and then Fe(OH)2 is dominant. The same behavior is noticed with aluminum electrodes. Therefore, the removal efficiency will be changed according to the type of the dominant specie of coagulants [70]. In conclusion, the pH of the solutions being treated by the EC units plays a significant role in the deterioration of the species and charge of the fresh coagulants, which in turn significantly determines the removal efficiency.

3.2. Role of Electrode Material

Selecting and employing feasible electrode material is essential to maximize the efficiency of the EC process. Besides conventional electrode materials, including Fe, Al, Cu, Mg, Sr and Zn, a few other materials, such as graphite and lead oxides, may seldom be used in the EC process as insoluble anodes. Despite the efficiency, easy availability and cost-effectiveness of conventional electrode materials, they have high oxygen evolution potential characteristics. Therefore, many materials were utilized in the EC method to remove pollutants. For example, Safwat [71] investigated the efficiency of zinc (Zn) and titanium (Ti) electrodes for the removal of oils, total dissolved solids (TDS) and COD from solutions at various experimental conditions (current density, spaces between electrodes and time). The results showed that the best oil removal efficiency was 14% by the Zn electrode and 19% by the Ti electrode, and the removal of COD was 50% and 46% by Zn and Ti electrodes, respectively. TDS removal efficiencies were 9% and 19% for Ti and Zn electrodes, respectively. These results showed that Zn and Ti electrodes could remove pollutants when used in the EC process. Hafez, et al. [72] used three different materials, Al, Fe and Zn, as electrodes to remove hardness and dissolved silica from wastewater under different experimental conditions. The obtained results showed the Al electrodes were better than the Fe and Zn electrodes, because Al removed 55.36% and 99.54% of hardness and dissolved silica. However, the low cost, better dissolution and availability of Fe and Al electrodes made them favorable in wastewater treatment practices. Reátegui-Romero, et al. [73] used iron electrodes to remove several metals (Fe, Mn, Cu, Zn and Cd) at pH between 6.6 to 8, a current density of 22.35 A/m2, and an application time of 45 min. The results showed the removal efficiency of Fe and Mn at 99.17% and 99.97%, respectively, whereas the removal efficiency of Cu, Zn and Cd was about 99.99%. In another study, Khosa et al., 2013 [74] used aluminum and iron electrodes to remove Ni, Pb and Cd. The results showed the Fe electrodes achieved a better removal; the Al electrodes removed 92.3%, 91.5% and 89.4% for Ni, Pb and Cd, respectively, while the Fe electrodes removed 99.0%, 98.3% and 96.8% of Ni, Pb and Cd, respectively.
The effects of the electrode materials on the performance of the EC method are related to many factors. For example, the chemical and physical properties of electrode materials are different in terms of ions’ size, the ions’ migration speed in solution, oxidation potential, ion-OH bond polarity and structure of their hydroxide compounds. Therefore, the adsorption capacity of the coagulants is different from one material to another [71,72,75]. The main conclusion of this short literature review is that the material of electrodes determines the chemical composition and adsorption capacity of the coagulants and the size of flocs, which determines the removal of the EC method.

3.3. Role of Current Density

The influence of the applied current density in the EC process is quite noteworthy, as it controls both bubble generation rate and coagulant dosage. The latter two parameters are very important because, firstly, the bubble generation rate determines the removal paths. Increasing the applied current increases the number of the generated bubbles, which results in the predomination of the floatation path in the separation of pollutants, while the low current densities lead to a low production of bubbles that makes the sedimentation path the predominant separation path. Additionally, the high current density increases the anodic dissolution that increases the formation of hydroxo-cationic complexes, thereby improving the removal efficiency (Liu et al., 2010). A considerable number of studies investigate the effects of the current density on the removal of pollutants by the EC units. For example, Manilal et al. [76] applied the EC method to remove oil and chemical oxygen demand (COD) from water, and the experiments were carried out using mild steel electrodes under different conditions, including the current density (between 0.25 to 0.75 A/dm2). The results showed the current density significantly affects the removal of the pollutant; it was noticed the removal of oil and COD increased from about 55% to about 90% when the current density increased from 0.25 to 0.75 A/dm2. The authors explained this increase in the removal of oil and COD with the current density by the rise in the production of the coagulants [Fe(OH)3 ions] that, in turn, produce more precipitates and consequently improve the removal of pollutants. A similar trend was noticed by Mena, et al. [77] during the removal of fluoride from water using aluminum electrodes under different current densities (in the range between 3 and 15 mA/cm2). The results showed that increasing the current density from 3 and 15 mA/cm2 considerably improved the removal of the fluoride from about 60% to about 80%, respectively. The authors also explained the increase in the removal of fluoride with the current density by the increase in the production of the aluminum ions, which improves the removal efficiency. The same effect of the applied current density on the removal of heavy metals (arsenic) from the water was noticed by Müller, et al. [78]. The results showed that the residual concentration of arsenic decreased from about 750 to 10 µg/L when the current density increased from 1.72 to 0.49 mA/cm2 after 180 min. The given reason for the increment in arsenic removal with the current density by the increase in the production of Fe(OH)3 ions that improve the removal of arsenic. The conclusion from these studies is that the effect of the current density on the efficiency of the EC method is related to the production rate of the coagulants from the anodes, which speeds up the removal of pollutants.

3.4. Operational Time

Operational time is another key parameter in EC, as the coagulant concentration generated due to electrolysis is time-bound and directly dependent on the electric charge added per unit volume. With longer durations, more hydroxyl radicals are generated, resulting in more availability of metallic polymer species in the form of hydroxides as adsorbents, thus increasing the removal efficiency [79,80]. However, the removal efficiency of pollutants increases with electrolysis time until it reaches saturation at an optimum condition [59]. This operation time is well associated with current density and thus needs to be optimized, as the removal efficiency desired depends upon the availability of coagulant floc for the target removal of pollutants. The effect of the operating time on the performance of the EC units was widely investigated in the literature; for example, Balarak, et al. [81] used a bench-scale EC unit to remove amoxicillin from water under different operating parameters and noticed increasing the treatment time significantly improves the removal of amoxicillin. The results showed that increasing the treatment time from 10 to 100 min increased the removal of amoxicillin by about 58%. Similar results were noticed in the removal of other pollutants, such as total organic carbon [82], orange II dye [83], Cr [84] and COD [85]. The given reason for the improvement of pollutant removal with the increase in the treatment time is more coagulants will be produced over time, which improves the removal of the targeted pollutants [81,86].
It can be concluded from the reviewed studies that it is essential to provide an adequate period for the EC treatment to generate coagulants, hence, to complete the coagulation of the pollutants.

3.5. Initial Concentration

The amount of metal ions present in an aqueous solution is another important parameter in the EC method. Several studies have been performed in the past to inspect its influence on the performance of the EC method [87,88]. For example, the effect of initial arsenic concentration on its removal efficiency was investigated by Chen, et al. [89] by preparing solutions with an initial concentration of zinc ranging from 50–2000 mg/L at a pH of 2.9–7.4 and a current density of 8.3 mA/cm2. It was observed that higher removal efficiency was achieved at lower initial concentration values, and it was found the residual concentration of zinc in the solution increased from 0 to about 1250 mg/L when the initial concentration of zinc increased from 50 to 2000 mg/L, respectively. Moreover, the removal rate was observed initially to be quite high in the early 10 min of the process; however, the rate declined afterwards. Moreover, it was found that amount of coagulant generated was almost the same having no effect of varying the initial concentration, due to which the generated coagulant was insufficient for the complete removal of metal ions at higher initial concentrations. Thereby, with a higher initial concentration of the ions present, there is more power consumption to attain the complete removal of target metal ions. These findings agree with another study carried out in treating Cr(VI) with initial concentrations of 329.55 to 1170.45 ppm using an EC unit supplied with iron electrodes [90]. The results showed increasing the initial concentration of Cr(VI) from 329.55 to 1170.45 ppm decreased the removal efficiency from 99.9% to 84.6%, respectively. The reviewed studies indicated that a decrease in the efficiency of the EC units with the increase of the concentration of the pollutant is due to insufficient production of the coagulant to remove the high concentrations, but this problem can be solved by increasing either the treatment time or the applied current density, which allows the anodes to produce more coagulants and thereby complete the removal of the high concentrations of the pollutants. The conclusion here is the decrease in the efficiency of the EC units with the increase in the concentration of the pollutant is not related to the performance of the EC reactor; it is related directly to the production rate of the coagulants. Therefore, this problem can be overcome by applying high current density to accelerate the dissolving of the anodes or by increasing the treatment time.

3.6. Electrode Arrangement

Typical arrangement of electrodes inside an EC reactor may be either bipolar or monopolar, such arrangements have been tried in various studies carried out in the past as shown in Figure 2. Several studies have comparatively reported the removal efficiency and operating cost from the monopolar and bipolar arrangement [91,92,93]. For instance, a group of researchers (Khaled et al. [94]) employed both monopolar and bipolar configurations using 4 aluminum electrodes for the removal of cadmium, with results showing rapid removal of up to 95% taking just 30 min, in comparison to when only two electrodes were used, where it took 45 min for almost the same performance. The removal efficiency obtained from the bipolar configuration was 92%, whereas 87% was obtained from the monopolar configuration. More often, effective treatment costs are incurred, and a high removal efficiency rate makes bipolar configuration more favorable. In another study, the removal efficiency of 99.84% and 89.58% for Cr (III) from bipolar and monopolar configuration after 20 min of EC at a current density of 25 mA/cm2 was reported [95]. In an earlier alternate study, the same removal efficiency from monopolar and bipolar configuration was found but reported bipolar configuration as energy and cost-effective [63].

3.7. Electrode Spacing

Spacing between electrodes is considered a vital parameter in the design of the reactor for the efficacy of the EC process, because it determines the energy consumption and reactor size [96]. Generally, the removal efficiency increases with the increase in electrode spacing until it reaches the optimum distance; then, the travel time of ions between electrodes increases resulting in reduced removal efficiency. Additionally, the literature shows that the potential developed in the EC units due to internal resistance (IR), is directly proportional to electrode distance; as the electrode spacing increases, the IR drop increases [75]. This drop in the IR increases power consumption and decreases the efficiency of the EC units. Therefore, optimizing the distance between electrodes is vital in the EC units [97]. Hence, most literature included the distance between electrodes as a crucial parameter. For example, several experiments were conducted by Mohammadi et al. [98] to measure the effects of different spaces between electrodes (3, 5, 6 and 7 cm) on the removal of nitrogen using an EC unit with iron and aluminum electrodes. It was found that closely spaced electrodes achieved higher removal efficiency; increasing the distances between electrodes from 3 to 7 cm decreased the removal of nitrogen from about 53 % to 20% and increased the power consumption from 4.75 to 19.01 KWh/kg of nitrogen, respectively. The authors explained the decrease in the performance of the EC unit by the fact that short distances decrease the IR drop, which facilitates the travel of ions between electrodes and minimizes power consumption. The relationship between the distance between electrodes and the IR is shown by the following equation [99]:
IR = Cuurent   A × Distnace   beween   electodes   m Effective   anode   area   m 2 × Specific   conductivity   10 3   mS / m  
The same results were noticed by other researchers, such as Naje et al. [100], Maghanga et al. [99], and Liu, et al. [101], and they justified the drop in the removal efficiency with the increase in the distance between electrodes by the reduction in the IR and the long travel time for the ions between the electrodes. The conclusion here is that the distance between electrodes must be optimized not because it affects the removal efficiency but because it also affects the power consumption, which determines the operating cost of the EC units.

3.8. Agitation

To sustain a uniform distribution of coagulants and pollutants inside an EC cell, stirring or agitation at a particular intensity is generally employed [102]. Obviously, by agitating the solution, the velocity attained by generated ions due to anode oxidation increases accordingly, which enhances the contact between the pollutants and the coagulants and thereby improves the removal efficiency. However, increasing the stirring speed more than the optimum value leads to breaking the fresh flocs, negatively affecting the removal efficiency [94,103]. The influence of agitation speed on the EC process has been investigated by numerous researchers in the past. For example, Khaled et al. [94] demonstrated the positive outcome of agitation on cadmium removal from wastewater by EC in which observations were taken at 0, 300, 450, and 600 rpm and found moderate agitation most effective. In this investigation, the maximum attainable removal efficiency was 95% and 99.5%, achieved after 30 min and 60 min, respectively. A slight decrease was noticed in the removal efficiency was observed at higher speeds, 450 rpm and 600 rpm, which could be due to the breaking of flocks and excessive agitation. Similarly, Bao et al. [103] investigated the effect of stirring speed on the removal of three types of perfluoro-alkane-sulfonic acids (PFBS, PFHxS and PFOS) from groundwater using an EC unit having Al and Zn electrodes. The experiments were conducted at stirring speeds of 0 to 1000 rpm, and results showed that a stirring speed of 400 rpm achieved the best removal of PFBS, PFHxS and PFOS (87.4%, 95.6% and 100%, respectively). The decrease in the removal of these pollutants at high stirring speeds was due to the development of a high shear force within the solution that destroys the hydroxide flocs, while the low efficiency at low stirring speeds is attributed to the insufficient contact between coagulants and pollutants (insufficient mixing). The same trend was found by Liu et al. [104] who reported that increasing the stirring speed more than the optimum value (500 rpm) decreases the removal efficiency of perfluoro-octanoic acid from water, and explained this reduction in the removal efficiency by the prevention of adsorption of the pollutant by the fresh coagulants. It can be concluded here that low or high stirring speeds are not useful for the EC treatment, and these effects can be summarized as follows:
  • High stirring speeds prevent the adsorption of pollutants on the fresh flocs and also break down the aggregates, which decreases the removal efficiency;
  • Low stirring speeds do not create a uniform distribution of pollutants and coagulants in the solution, which also decreases the efficiency of the EC unit.

4. Kinetic Models Applied in EC

Although it has quite a complex mechanism, including various physio-chemical reactions, the adsorption is only one step in the whole electrocoagulation (EC) process. The complexes produced during EC as a result of oxidation in the form of metal hydroxide act as an adsorbent for pollutants in the water and wastewater. These adsorption and complexation steps in the EC process can be modelled using suitable models, such as the first-order, second-order, Dubinin–Radushkevich model, Elovich model, Temkin model, Pseudo-first-order or Pseudo-second-order [52,105]. Each of these models has its peculiarities and significances. For instance, the Pseudo-first-order model describes the adsorbate uptake and is based on solid adsorption capacity, while the Pseudo-second-order model undertakes chemisorption involving the exchange of electrons between metal ions and OH. For instance, Yılmaz et al. [106] carried out a statistical analysis, along with an empirical model, to explain the boron removal from wastewater stream employing aluminum electrodes in EC. They reported that the reaction rate followed the pseudo-second-order model quite well.
The pseudo-first-order model is as follows:
dq dt = k 1   ( q e q )
where k1 is the first-order rate constant, q is the adsorption capacity at any time “t“ and qe is the adsorption capacity on attaining equilibrium. Integrating and linearizing this equation, the resulting Equation is as follows:
ln(qe − q) = ln(qe) − k1t
ln(qe− q) is plotted against “t” and linear regression is carried out to obtain the best fit. The values of qe and k1 were acquired from the slope and intercept of the plots accordingly.
The Pseudo-second-order model is given as follows:
dq dt = k 2 ( q e q ) 2
where k2 is the second-order rate constant. The following equation can be obtained by integrating and rearranging Equation (12):
t q = 1 k 2 q e 2 + t q e
The graph of “ t q ” vs. ”t” is plotted, and linear regression is carried out to obtain the equation. qe and k2 values are depicted from the slope and intercept of the plots accordingly.
Several researchers have applied kinetic modelling in the adsorption of heavy metals in the EC process. Xi, et al. [107] applied both first and second-order models in the removal of cadmium using EC and found better values of correlation coefficients in the case of second-order model plots. Similarly, in another research carried out on the removal of mercury, lead, and nickel using EC, the second-order model was found to be more suitable in explaining the adsorption of all three metal ions over Mg(OH)2 by employing magnesium anode [108]. However, in another investigation carried out employing iron electrodes in EC for heavy metal removal, the Pseudo-second-order model was found to fit well [52]. In the same manner, several other researchers have explored the suitability of these models in their studies, a few of which are summarized in Table 4. Thus it may be concluded that various models are used to explain the adsorption step of heavy metal in EC. This may be attributed because the interaction between pollutants and coagulant ions depends on their nature, properties, and operating conditions.

5. Cost Analysis

Cost information and estimation are available in very few researches employing EC. Considerable work at the laboratory and pilot scale had been conducted using EC processes for treating metal-containing wastewater, and it has been speculated that the cost is quite competitive with traditional chemical precipitation and other techniques. The main constraint in the large-scale applicability of the EC process is its operating cost [116]. For wider applicability in treating various industrial effluents, it is necessary to make EC a low-cost technique. Increasing energy prices is an important issue that discourages the applicability of the EC process on a field scale [117]. Moreover, electrodes get sacrificed in dissipating coagulants and thus need to be replaced from time to time, which adds up to the overall operational cost of the process. Therefore, optimization of the EC process is a necessary step in making it cost-effective before any practical use, thereby minimizing energy consumption and attaining higher efficiency. For instance, Khan et al. [28] reduced the overall cost of the EC process to a very low amount in treating per g of Cr(VI) by multivariate optimization of process variables. In another research, response surface methodology (RSM), a mathematical approach, was used to optimize and analyze the effect of the process-dependent variables, affecting the treatment efficiency and thus making the process cost-effective [118]. They attained higher efficiencies at lower operating costs, suggesting the successful applicability of RSM and the need for multivariate optimization as a whole in achieving the objectives. A good illustration of cost variation for the EC process was performed by Shaker et al. [119] on nickel removal at different current densities (5, 10, and 15 mA/cm2). They found that the least EC cost was around 0.7 USD/m3 when the current density was five mA/cm2 with an initial pH of 9.2. From an economic aspect, a current density of 5 mA/cm2 instead of 10 mA/cm2 is more feasible, because the difference in the removal efficiency is insignificant.
Few comparative studies on optimizing the removal of heavy metal using EC have been tabulated, indicating the energy consumption as well as operating cost (see Table 5). Another potential advantage of the EC process is the lesser production of sludge which is generally easier to dewater that might be usefully recovered.

6. Conclusion Remarks and Future Perspectives

The EC method has gained wider acceptance as an alternative to conventional wastewater treatment methods, although its potential needs further exploration. Simple apparatus, expedient operation and no use of chemical substances for settling and floc formation are the main advantages of this process. EC had been successfully applied with a variable degree of treatment achieved for various types of water and wastewater streams. The current density, electrolysis time and electrode type must be carefully considered for each application. It is found that this treatment process has been successfully used to treat different types of heavy metal ions, such as Ni, Cu, Cr, Cd, etc. Optimization of this process not only enhances efficiency, but also makes it possible for its use at a larger scale at an affordable cost. The EC process has proven to be a practical, cost-effective technology for removing most pollutants. However, in the future, more work needs to be done in process optimization, modelling and scaling up this technique to prove its consistency for efficient and large-scale wastewater treatment. Multivariate optimization techniques, such as RSM, have made the applicability of this technique wider. In addition, sustainable end-use of sludge generated is essential, as it has recently become an attractive research area. Beyond that, we may improve and enhance the efficiency of the EC process by combining it with other conventional techniques or using other methods sequentially as pre-treatment or as a polishing step wherever required.

Author Contributions

Conceptualization, S.U.K. and M.K.; data curation, S.U.K., M.K. and K.H.; formal analysis, S.U.K. and M.K.; investigation, S.U.K., M.K. and M.H.J.; methodology, S.U.K., M.K., F.B., I.H.F. and F.B.; project administration, S.U.K.; resources, S.U.K.; software, S.U.K. and M.K.K; supervision, S.U.K.; validation, S.U.K. and M.M.; visualization, S.U.K., K.H. and M.K.K; writing—original draft, S.U.K. and M.K.K; writing—review and editing, S.U.K., M.K.K, K.H., M.H.J., M.M., F.B., I.H.F. and F.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yan, C.; Qu, Z.; Wang, J.; Cao, L.; Han, Q. Microalgal bioremediation of heavy metal pollution in water: Recent advances, challenges, and prospects. Chemosphere 2022, 286, 131870. [Google Scholar] [CrossRef] [PubMed]
  2. Al-Hashimi, O.; Hashim, K.; Loffill, E.; Nakouti, I.; Faisal, A.A.; Čebašek, T.M. Kinetic and Equilibrium Isotherm Studies for the Removal of Tetracycline from Aqueous Solution Using Engineered Sand Modified with Calcium Ferric Oxides. Environments 2023, 10, 7. [Google Scholar] [CrossRef]
  3. Smal, H.; Ligęza, S.; Wójcikowska-Kapusta, A.; Baran, S.; Urban, D.; Obroślak, R.; Pawłowski, A. Spatial distribution and risk assessment of heavy metals in bottom sediments of two small dam reservoirs (south-east Poland). Arch. Environ. Prot. 2015, 41, 67–80. [Google Scholar] [CrossRef]
  4. Omran, I.I.; Al-Saati, N.H.; Al-Saati, H.H.; Hashim, K.S.; Al-Saati, Z.N. Sustainability assessment of wastewater treatment techniques in urban areas of Iraq using multi-criteria decision analysis (MCDA). Water Pract. Technol. 2021, 16, 648–660. [Google Scholar] [CrossRef]
  5. Khan, N.A.; Khan, S.U.; Islam, D.T.; Ahmed, S.; Farooqi, I.H.; Isa, M.H.; Hussain, A.; Changani, F.; Dhingra, A. Performance evaluation of column-SBR in paper and pulp wastewater treatment: Optimization and bio-kinetics. Desalin Water Treat. 2019, 156, 204–219. [Google Scholar] [CrossRef] [Green Version]
  6. Khan, S.U.; Noor, A.; Farooqi, I.H. GIS application for groundwater management and quality mapping in rural areas of District Agra, India. Int. J. Water Resour. Arid. Environ. 2015, 4, 89–96. [Google Scholar]
  7. Puttaiah, E.T.; Kiran, B.R. Heavy metal transport in a sewage fed lake of Karnataka, India. In Proceedings of the TAAL: The World Lake Conference, Jaipur, India, 28 October–2 November 2007; pp. 347–354. [Google Scholar]
  8. Mukherjee, I.; Singh, U.K.; Singh, R.P. An overview on heavy metal contamination of water system and sustainable approach for remediation. In Water Pollution and Management Practices; Springer: Singapore, 2021; pp. 255–277. [Google Scholar]
  9. Drasch, G.; Schöpfer, J.; Schrauzer, G.N. Selenium/cadmium ratios in human prostates. Biol. Trace Elem. Res. 2005, 103, 103–107. [Google Scholar] [CrossRef]
  10. Kumar, L.; Chugh, M.; Kumar, S.; Kumar, K.; Sharma, J.; Bharadvaja, N. Remediation of petrorefinery wastewater contaminants: A review on physicochemical and bioremediation strategies. Process Saf. Environ. Prot. 2022, 1, 362–375. [Google Scholar] [CrossRef]
  11. Simeonov, V.; Stratis, J.; Samara, C.; Zachariadis, G.; Voutsa, D.; Anthemidis, A.; Sofoniou, M.; Kouimtzis, T. Assessment of the surface water quality in Northern Greece. Water Res. 2003, 37, 4119–4124. [Google Scholar] [CrossRef]
  12. Tamasi, G.; Cini, R. Heavy metals in drinking waters from Mount Amiata (Tuscany, Italy). Possible risks from arsenic for public health in the Province of Siena. Sci. Total Environ. 2004, 327, 41–51. [Google Scholar] [CrossRef]
  13. Demirak, A.; Yilmaz, F.; Tuna, A.L.; Ozdemir, N. Heavy metals in water, sediment and tissues of Leuciscus cephalus from a stream in southwestern Turkey. Chemosphere 2006, 63, 1451–1458. [Google Scholar] [CrossRef]
  14. Buschmann, J.; Berg, M.; Stengel, C.; Winkel, L.; Sampson, M.L.; Trang, P.T.K.; Viet, P.H. Contamination of drinking water resources in the Mekong delta floodplains: Arsenic and other trace metals pose serious health risks to population. Environ. Int. 2008, 34, 756–764. [Google Scholar] [CrossRef]
  15. Karavoltsos, S.; Sakellari, A.; Mihopoulos, N.; Dassenakis, M.; Scoullos, M.J. Evaluation of the quality of drinking water in regions of Greece. Desalination 2008, 224, 317–329. [Google Scholar] [CrossRef]
  16. El-Harouny, M.A.; El-Dakroory, S.A.; Attalla, S.M.; Hasan, N.A.; Hegazy, R. Chemical quality of tap water versus bottled water: Evaluation of some heavy metals and elements content of drinking water in Dakahlia governorate-Egypt. Mansoura J. Forensic Med. Clin. Toxicol. 2008, 16, 1–15. [Google Scholar] [CrossRef]
  17. Chakraborti, D.; Rahman, M.M.; Das, B.; Murrill, M.; Dey, S.; Mukherjee, S.C.; Dhar, R.K.; Biswas, B.K.; Chowdhury, U.K.; Roy, S. Status of groundwater arsenic contamination in Bangladesh: A 14-year study report. Water Res. 2010, 44, 5789–5802. [Google Scholar] [CrossRef]
  18. Ahmad, M.; Islam, S.; Rahman, S.; Haque, M.; Islam, M. Heavy metals in water, sediment and some fishes of Buriganga River, Bangladesh. Int. J. Environ. Res. 2010, 4, 321–332. [Google Scholar]
  19. Khdary, N.H.; Gassim, A.E. The distribution and accretion of some heavy metals in Makkah Wells. J. Water Resour. Prot. 2014, 6, 48906. [Google Scholar] [CrossRef] [Green Version]
  20. Mandour, R.; Azab, Y. The prospective toxic effects of some heavy metals overload in surface drinking water of Dakahlia Governorate, Egypt. Int. J. Occup. Environ. Med. 2011, 2, 245–253. [Google Scholar]
  21. Islam, M.; Karim, M.; Higuchi, T.; Sakakibara, H.; Sekine, M. Comparison of the trace metal concentration of drinking water supply options in southwest coastal areas of Bangladesh. Appl. Water Sci. 2014, 4, 183–191. [Google Scholar] [CrossRef] [Green Version]
  22. Ojekunle, O.Z.; Ojekunle, O.V.; Adeyemi, A.A.; Taiwo, A.G.; Sangowusi, O.R.; Taiwo, A.M.; Adekitan, A.A. Evaluation of surface water quality indices and ecological risk assessment for heavy metals in scrap yard neighbourhood. SpringerPlus 2016, 5, 1–16. [Google Scholar] [CrossRef] [Green Version]
  23. Boateng, T.K.; Opoku, F.; Akoto, O. Heavy metal contamination assessment of groundwater quality: A case study of Oti landfill site, Kumasi. Appl. Water Sci. 2019, 9, 1–15. [Google Scholar] [CrossRef] [Green Version]
  24. de Oliveira, M.T.; Torres, I.M.S.; Ruggeri, H.; Scalize, P.; Albuquerque, A.; Gil, E.d.S. Application of electrocoagulation with a new steel-swarf-based electrode for the removal of heavy metals and total coliforms from sanitary landfill leachate. Appl. Sci. 2021, 11, 5009. [Google Scholar] [CrossRef]
  25. Stylianou, M.; Montel, E.; Zissimos, A.; Christoforou, I.; Dermentzis, K.; Agapiou, A. Removal of toxic metals and anions from acid mine drainage (AMD) by electrocoagulation: The case of North Mathiatis open cast mine. Sustain. Chem. Pharm. 2022, 29, 100737. [Google Scholar] [CrossRef]
  26. Beiramzadeh, Z.; Baqersad, M.; Aghababaei, M. Application of the response surface methodology (RSM) in heavy metal removal from real power plant wastewater using electrocoagulation. Eur. J. Environ. Civ. Eng. 2022, 26, 1–20. [Google Scholar] [CrossRef]
  27. Xu, R.; Yang, Z.; Niu, Y.; Xu, D.; Wang, J.; Han, J.; Wang, H. Removal of microplastics and attached heavy metals from secondary effluent of wastewater treatment plant using interpenetrating bipolar plate electrocoagulation. Sep. Purif. Technol. 2022, 290, 120905. [Google Scholar] [CrossRef]
  28. Khan, N.A.; Khan, S.U.; Ahmed, S.; Farooqi, I.H.; Dhingra, A.; Hussain, A.; Changani, F. Applications of nanotechnology in water and wastewater treatment: A review. Asian J. Water Environ. Pollut. 2019, 16, 81–86. [Google Scholar] [CrossRef]
  29. Khadem, M.; Husni Ibrahim, A.; Mokashi, I.; Hasan Fahmi, A.; Noeman Taqui, S.; Mohanavel, V.; Hossain, N.; Baba Koki, I.; Elfasakhany, A.; Dhaif-Allah, M.A. Removal of heavy metals from wastewater using low-cost biochar prepared from jackfruit seed waste. Biomass Convers. Biorefinery 2022, 1–10. [Google Scholar] [CrossRef]
  30. Zaidi, R.; Khan, S.U.; Azam, A.; Farooqi, I.H. A study on effective adsorption of lead from an aqueous solution using copper oxide nanoparticles. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1058, 012074. [Google Scholar] [CrossRef]
  31. Kumari, S.; Kumar, R.N. River water treatment using electrocoagulation for removal of acetaminophen and natural organic matter. Chemosphere 2021, 273, 128571. [Google Scholar] [CrossRef]
  32. Hendaoui, K.; Trabelsi-Ayadi, M.; Ayari, F. Optimization and mechanisms analysis of indigo dye removal using continuous electrocoagulation. Chin. J. Chem. Eng. 2021, 29, 242–252. [Google Scholar] [CrossRef]
  33. Alam, R.; Sheob, M.; Saeed, B.; Khan, S.U.; Shirinkar, M.; Frontistis, Z.; Basheer, F.; Farooqi, I.H. Use of electrocoagulation for treatment of pharmaceutical compounds in water/wastewater: A review exploring opportunities and challenges. Water 2021, 13, 2105. [Google Scholar] [CrossRef]
  34. Hashim, K.S.; AlKhaddar, R.; Shaw, A.; Kot, P.; Al-Jumeily, D.; Alwash, R.; Aljefery, M.H. Electrocoagulation as an eco-friendly River water treatment method. In Advances in Water Resources Engineering and Management; Springer: Singapore, 2020; pp. 219–235. [Google Scholar] [CrossRef]
  35. Taqui, S.N.; Cs, M.; Goodarzi, M.S.; Elkotb, M.A.; Khatoon, B.A.; Soudagar, M.E.M.; Koki, I.B.; Elfasakhany, A.; Khalifa, A.S.; Ali, M.A. Sustainable adsorption method for the remediation of crystal violet dye using nutraceutical industrial fenugreek seed spent. Appl. Sci. 2021, 11, 7635. [Google Scholar] [CrossRef]
  36. Mahvi, A.; Malakootian, M.; Heidari, M. Comparison of polyaluminum silicate chloride and electrocoagulation process, in natural organic matter removal from surface water in Ghochan, Iran. J. Water Chem. Technol. 2011, 33, 377–385. [Google Scholar] [CrossRef] [Green Version]
  37. Hashim, K.S.; Ali, S.S.M.; AlRifaie, J.K.; Kot, P.; Shaw, A.; Al Khaddar, R.; Idowu, I.; Gkantou, M. Escherichia coli inactivation using a hybrid ultrasonic–electrocoagulation reactor. Chemosphere 2020, 247, 125868–125875. [Google Scholar] [CrossRef]
  38. Ölmez, T. The optimization of Cr (VI) reduction and removal by electrocoagulation using response surface methodology. J. Hazard. Mater. 2009, 162, 1371–1378. [Google Scholar] [CrossRef]
  39. Gholami Borujeni, F.; Mahvi, A.H.; Nejatzadeh-Barandoozi, F. Removal of heavy metal ions from aqueous solution by application of low cost materials. Fresen. Environ. Bull. 2013, 22, 655–658. [Google Scholar]
  40. Khan, S.U.; Zaidi, R.; Hassan, S.Z.; Farooqi, I.; Azam, A. Application of Fe-Cu binary oxide nanoparticles for the removal of hexavalent chromium from aqueous solution. Water Sci. Technol. 2016, 74, 165–175. [Google Scholar] [CrossRef]
  41. Chen, Q.; Luo, Z.; Hills, C.; Xue, G.; Tyrer, M. Precipitation of heavy metals from wastewater using simulated flue gas: Sequent additions of fly ash, lime and carbon dioxide. Water Res. 2009, 43, 2605–2614. [Google Scholar] [CrossRef]
  42. Qdais, H.A.; Moussa, H. Removal of heavy metals from wastewater by membrane processes: A comparative study. Desalination 2004, 164, 105–110. [Google Scholar]
  43. Tan, J.-j.; Huang, Y.; Wu, Z.-q.; Chen, X. Ion exchange resin on treatment of copper and nickel wastewater. In Proceedings of the IOP Conference Series: Earth and Environmental Science, Zvenigorod, Russia, 4–7 September 2017; p. 012122. [Google Scholar]
  44. Akbal, F.; Camcı, S. Copper, chromium and nickel removal from metal plating wastewater by electrocoagulation. Desalination 2011, 269, 214–222. [Google Scholar] [CrossRef]
  45. Goren, A.; Kobya, M. Arsenic removal from groundwater using an aerated electrocoagulation reactor with 3D Al electrodes in the presence of anions. Chemosphere 2021, 263, 128253. [Google Scholar] [CrossRef] [PubMed]
  46. Mamelkina, M.A.; Tuunila, R.; Sillänpää, M.; Häkkinen, A. Systematic study on sulfate removal from mining waters by electrocoagulation. Sep. Purif. Technol. 2019, 216, 43–50. [Google Scholar] [CrossRef]
  47. Asaithambi, P.; Govindarajan, R.; Yesuf, M.B.; Selvakumar, P.; Alemayehu, E. Investigation of direct and alternating current–electrocoagulation process for the treatment of distillery industrial effluent: Studies on operating parameters. J. Environ. Chem. Eng. 2021, 9, 104811. [Google Scholar] [CrossRef]
  48. Ano, J.; Briton, B.G.H.; Kouassi, K.E.; Adouby, K. Nitrate removal by electrocoagulation process using experimental design methodology: A techno-economic optimization. J. Environ. Chem. Eng. 2020, 8, 104292. [Google Scholar] [CrossRef]
  49. Núñez, J.; Yeber, M.; Cisternas, N.; Thibaut, R.; Medina, P.; Carrasco, C. Application of electrocoagulation for the efficient pollutants removal to reuse the treated wastewater in the dyeing process of the textile industry. J. Hazard. Mater. 2019, 371, 705–711. [Google Scholar] [CrossRef] [PubMed]
  50. Jallouli, S.; Wali, A.; Buonerba, A.; Zarra, T.; Belgiorno, V.; Naddeo, V.; Ksibi, M. Efficient and sustainable treatment of tannery wastewater by a sequential electrocoagulation-UV photolytic process. J. Water Process Eng. 2020, 38, 101642. [Google Scholar] [CrossRef]
  51. Dimoglo, A.; Sevim-Elibol, P.; Dinç, Ö.; Gökmen, K.; Erdoğan, H. Electrocoagulation/electroflotation as a combined process for the laundry wastewater purification and reuse. J. Water Process Eng. 2019, 31, 100877. [Google Scholar] [CrossRef]
  52. Al-Shannag, M.; Al-Qodah, Z.; Bani-Melhem, K.; Qtaishat, M.R.; Alkasrawi, M. Heavy metal ions removal from metal plating wastewater using electrocoagulation: Kinetic study and process performance. Chem. Eng. J. 2015, 260, 749–756. [Google Scholar] [CrossRef]
  53. Lu, J.; Fan, R.; Wu, H.; Zhang, W.; Li, J.; Zhang, X.; Sun, H.; Liu, D. Simultaneous removal of Cr (VI) and Cu (II) from acid wastewater by electrocoagulation using sacrificial metal anodes. J. Mol. Liq. 2022, 359, 119276. [Google Scholar] [CrossRef]
  54. Foudhaili, T.; Lefebvre, O.; Coudert, L.; Neculita, C.M. Sulfate removal from mine drainage by electrocoagulation as a stand-alone treatment or polishing step. Miner. Eng. 2020, 152, 106337. [Google Scholar] [CrossRef]
  55. Mehri, M.; Fallah, N.; Nasernejad, B. Mechanisms of heavy metal and oil removal from synthetic saline oilfield produced water by electrocoagulation. NPJ Clean Water 2021, 4, 1–14. [Google Scholar] [CrossRef]
  56. Verma, S.K.; Khandegar, V.; Saroha, A.K. Removal of chromium from electroplating industry effluent using electrocoagulation. J. Hazard. Toxic Radioact. Waste 2013, 17, 146–152. [Google Scholar] [CrossRef]
  57. Dermentzis, K.; Valsamidou, E.; Lazaridou, A.; Kokkinos, N. Nickel removal from wastewater by electrocoagulation with aluminum electrodes. J. Eng. Sci. Technol. Rev. 2011, 4, 188–192. [Google Scholar] [CrossRef]
  58. Flores, O.J.; Nava, J.L.; Carreño, G. Arsenic removal from groundwater by electrocoagulation process in a filter-press-type FM01-LC reactor. Int. J. Electrochem. Sci 2014, 9, 6658–6667. [Google Scholar]
  59. Can, B.Z.; Boncukcuoglu, R.; Yilmaz, A.E.; Fil, B.A. Effect of some operational parameters on the arsenic removal by electrocoagulation using iron electrodes. J. Environ. Health Sci. Eng. 2014, 12, 1–10. [Google Scholar] [CrossRef]
  60. Daniel, R.; Himabindu, V.; Krupadam, R.; Anjaneyulu, Y. Removal of Mercury from Pharmaceutical Wastewaters Using Electrocoagulation: A Cleaner Technology Option. IUP J. Environ. Sci. 2010, 4, 7–15. [Google Scholar]
  61. Murthy, Z.; Parmar, S. Electrocoagulative treatment of mercury containing aqueous solutions. Water Sci. Technol. 2012, 65, 1468–1474. [Google Scholar] [CrossRef]
  62. Khan, S.U.; Mahtab, M.S.; Farooqi, I.H. Enhanced lead (II) removal with low energy consumption in an electrocoagulation column employing concentric electrodes: Process optimisation by RSM using CCD. Int. J. Environ. Anal. Chem. 2021, 1–18. [Google Scholar] [CrossRef]
  63. Kamaraj, R.; Ganesan, P.; Vasudevan, S. Removal of lead from aqueous solutions by electrocoagulation: Isotherm, kinetics and thermodynamic studies. Int. J. Environ. Sci. Technol. 2015, 12, 683–692. [Google Scholar] [CrossRef] [Green Version]
  64. Dermentzis, K.; Christoforidis, A.; Valsamidou, E. Removal of nickel, copper, zinc and chromium from synthetic and industrial wastewater by electrocoagulation. Int. J. Environ. Sci. 2011, 1, 697–710. [Google Scholar]
  65. Sasson, M.B.; Calmano, W.; Adin, A. Iron-oxidation processes in an electroflocculation (electrocoagulation) cell. J. Hazard. Mater. 2009, 171, 704–709. [Google Scholar] [CrossRef]
  66. Parga, J.R.; Cocke, D.L.; Valverde, V.; Gomes, J.A.; Kesmez, M.; Moreno, H.; Weir, M.; Mencer, D. Characterization of electrocoagulation for removal of chromium and arsenic. Chem. Eng. Technol. Ind. Chem.-Plant Equip.-Process Eng.-Biotechnol. 2005, 28, 605–612. [Google Scholar] [CrossRef]
  67. Al-Anbari, R.H.; Albaidani, J.; Alfatlawi, S.M.; Al-Hamdani, T.A. Removal of heavy metals from industrial water using electro-coagulation technique. In Proceedings of the Twelft International Water Technology Conference, Alexandria, Egypt, 27–30 March 2008. [Google Scholar]
  68. Kim, T.; Kim, T.-K.; Zoh, K.-D. 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. 2020, 33, 101109. [Google Scholar] [CrossRef]
  69. Xie, J.; Zhong, Y.; Yu, Y.; Wang, M.; Guo, Z. Green capturing of Ag from ultra-low concentration precious metal wastewater by electrodeposition assisted with electrocoagulation: Electrochemical behavior and floc characterization. Process Saf. Environ. Prot. 2022, 167, 592–600. [Google Scholar] [CrossRef]
  70. Barrera-Díaz, C.E.; Balderas-Hernández, P.; Bilyeu, B. Electrocoagulation: Fundamentals and prospectives. In Electrochemical Water and Wastewater Treatment; Elsevier: Amsterdam, The Netherlands, 2018; pp. 61–76. [Google Scholar]
  71. Safwat, S.M. Treatment of real printing wastewater using electrocoagulation process with titanium and zinc electrodes. J. Water Process Eng. 2020, 34, 101137. [Google Scholar] [CrossRef]
  72. Hafez, O.M.; Shoeib, M.A.; El-Khateeb, M.A.; Abdel-Shafy, H.I.; Youssef, A.O. Removal of scale forming species from cooling tower blowdown water by electrocoagulation using different electrodes. Chem. Eng. Res. Des. 2018, 136, 347–357. [Google Scholar] [CrossRef]
  73. Reátegui-Romero, W.; Flores-Del Pino, L.V.; Guerrero-Guevara, J.L.; Castro-Torres, J. Benefits of electrocoagulation in treatment of wastewater: Removal of Fe and Mn metals, oil and grease and cod: Three case studies. Int. J. Appl. Eng. Res. 2018, 13, 6450–6462. [Google Scholar]
  74. Khosa, M.K.; Jamal, M.A.; Hussain, A.; Muneer, M.; Zia, K.M.; Hafeez, S. Efficiency of Aluminum and Iron Electrodes for the Removal of Heavy Metals [(Ni (II), Pb (II), Cd (II)] by Electrocoagulation Method. J. Korean Chem. Soc. 2013, 57, 316–321. [Google Scholar] [CrossRef]
  75. Al-Qodah, Z.; Al-Shannag, M. Heavy metal ions removal from wastewater using electrocoagulation processes: A comprehensive review. Sep. Sci. Technol. 2017, 52, 2649–2676. [Google Scholar] [CrossRef]
  76. Manilal, A.; Soloman, P.; Basha, C.A. Removal of oil and grease from produced water using electrocoagulation. J. Hazard. Toxic Radioact. Waste 2020, 24, 04019023. [Google Scholar] [CrossRef]
  77. Mena, V.; Betancor-Abreu, A.; González, S.; Delgado, S.; Souto, R.; Santana, J. Fluoride removal from natural volcanic underground water by an electrocoagulation process: Parametric and cost evaluations. J. Environ. Manag. 2019, 246, 472–483. [Google Scholar] [CrossRef]
  78. Müller, D.; Stirn, C.N.; Maier, M.V. Arsenic Removal from Highly Contaminated Groundwater by Iron Electrocoagulation—Investigation of Process Parameters and Iron Dosage Calculation. Water 2021, 13, 687. [Google Scholar] [CrossRef]
  79. Asaithambi, P.; Aziz, A.R.A.; Daud, W.M.A.B.W. Integrated ozone—Electrocoagulation process for the removal of pollutant from industrial effluent: Optimization through response surface methodology. Chem. Eng. Process. Process Intensif. 2016, 105, 92–102. [Google Scholar] [CrossRef]
  80. Khan, S.U.; Islam, D.T.; Farooqi, I.H.; Ayub, S.; Basheer, F. Hexavalent chromium removal in an electrocoagulation column reactor: Process optimization using CCD, adsorption kinetics and pH modulated sludge formation. Process Saf. Environ. Prot. 2019, 122, 118–130. [Google Scholar] [CrossRef]
  81. Balarak, D.; Chandrika, K.; Attaolahi, M. Assessment of effective operational parameters on removal of amoxicillin from synthetic wastewater using electrocoagulation process. J. Pharm. Res. Int. 2019, 29, 8. [Google Scholar] [CrossRef] [Green Version]
  82. AlJaberi, F.Y. Removal of TOC from oily wastewater by electrocoagulation technology. Proc. IOP Conf. Ser. Mater. Sci. Eng. 2020, 928, 022024. [Google Scholar] [CrossRef]
  83. Signorelli, S.C.M.; Costa, J.M.; de Almeida Neto, A.F. Electrocoagulation-flotation for orange II dye removal: Kinetics, costs, and process variables effects. J. Environ. Chem. Eng. 2021, 9, 106157. [Google Scholar] [CrossRef]
  84. Patel, S.R.; Parikh, S.P. Chromium removal from industrial effluent by electrocoagulation: Operating cost and kinetic analysis. J. Environ. Treat. Tech. 2021, 9, 621–628. [Google Scholar]
  85. Alkurdi, S.S.; Abbar, A.H. Removal of COD from Petroleum refinery Wastewater by Electro-Coagulation Process Using SS/Al electrodes. Proc. IOP Conf. Ser. Mater. Sci. Eng. 2020, 870, 012052. [Google Scholar] [CrossRef]
  86. Danial, R.; Abdullah, L.C.; Sobri, S. Potential of copper electrodes in electrocoagulation process for glyphosate herbicide removal. Proc. MATEC Web Conf. 2017, 103, 06019. [Google Scholar] [CrossRef] [Green Version]
  87. Chen, M.; Dollar, O.; Shafer-Peltier, K.; Randtke, S.; Waseem, S.; Peltier, E. Boron removal by electrocoagulation: Removal mechanism, adsorption models and factors influencing removal. Water Res. 2020, 170, 115362. [Google Scholar] [CrossRef] [PubMed]
  88. Nyangi, M.J.; Chebude, Y.; Kilulya, K.F.; Andrew, M. Simultaneous removal of fluoride and arsenic from water by hybrid Al-Fe electrocoagulation: Process optimization through surface response method. Sep. Sci. Technol. 2021, 56, 2648–2658. [Google Scholar] [CrossRef]
  89. Chen, X.; Ren, P.; Li, T.; Trembly, J.P.; Liu, X. Zinc removal from model wastewater by electrocoagulation: Processing, kinetics and mechanism. Chem. Eng. J. 2018, 349, 358–367. [Google Scholar] [CrossRef]
  90. Genawi, N.M.; Ibrahim, M.H.; El-Naas, M.H.; Alshaik, A.E. Chromium removal from tannery wastewater by electrocoagulation: Optimization and sludge characterization. Water 2020, 12, 1374. [Google Scholar] [CrossRef]
  91. Othmani, A.; Kadier, A.; Singh, R.; Igwegb, C.A.; Bouzid, M.; Aquatar, M.O.; Khanday, W.A.; Bote, M.E.; Damiri, F.; Gökkuş, Ö. A comprehensive review on green perspectives of electrocoagulation integrated with advanced processes for effective pollutants removal from water environment. Environ. Res. 2022, 215, 114294. [Google Scholar] [CrossRef]
  92. Bagastyo, A. Electrocoagulation for drinking water treatment: A review. Proc. IOP Conf. Ser. Earth Environ. Sci. 2021, 623, 012016. [Google Scholar]
  93. Anwer, E.A.; Majeed, B.A.A. Different electrodes connections in electrocoagulation of synthetic blow down water of cooling tower. Iraqi J. Chem. Pet. Eng. 2020, 21, 1–7. [Google Scholar] [CrossRef] [Green Version]
  94. Khaled, B.; Wided, B.; Béchir, H.; Elimame, E.; Mouna, L.; Zied, T. Investigation of electrocoagulation reactor design parameters effect on the removal of cadmium from synthetic and phosphate industrial wastewater. Arab. J. Chem. 2019, 12, 1848–1859. [Google Scholar] [CrossRef]
  95. Al Aji, B. A Comparison Between monopolar and bipolar Electrode configurations on Removal of chromium from synthetic wastewater by Electrocoagulation Process. J. Res. Sci. Stud. 2016, 38, 525–537. [Google Scholar]
  96. Khandegar, V.; Saroha, A.K. Electrochemical treatment of effluent from small-scale dyeing unit. Indian Chem. Eng. 2013, 55, 112–120. [Google Scholar] [CrossRef]
  97. Salman, R.H.; Hassan, H.A.; Abed, K.M.; Al-Alawy, A.F.; Tuama, D.A.; Hussein, K.M.; Jabir, H.A. Removal of chromium ions from a real wastewater of leather industry using electrocoagulation and reverse osmosis processes. Proc. AIP Conf. Proc. 2020, 2213, 020186. [Google Scholar]
  98. Mohammadi, A.; Khadir, A.; Tehrani, R.M. Optimization of nitrogen removal from an anaerobic digester effluent by electrocoagulation process. J. Environ. Chem. Eng. 2019, 7, 103195. [Google Scholar] [CrossRef]
  99. Maghanga, J.; Segor, K.; Irina, J.; Tole, M. Effect of Process Parameters on the Electro Coagulation of Azo-Dye Wastewater in a Kenyan Textile Factory. IOSR J. Appl. Chem. 2017, 10, 1–7. [Google Scholar]
  100. Naje, A.S.; Chelliapan, S.; Zakaria, Z.; Abbas, S.A. Electrocoagulation using a rotated anode: A novel reactor design for textile wastewater treatment. J. Environ. Manag. 2016, 176, 34–44. [Google Scholar] [CrossRef]
  101. Liu, S.; Ye, X.; He, K.; Chen, Y.; Hu, Y. Simultaneous removal of Ni (II) and fluoride from a real flue gas desulfurization wastewater by electrocoagulation using Fe/C/Al electrode. J. Water Reuse Desalination 2017, 7, 288–297. [Google Scholar] [CrossRef] [Green Version]
  102. Naje, A.S.; Chelliapan, S.; Zakaria, Z.; Ajeel, M.A.; Alaba, P.A. A review of electrocoagulation technology for the treatment of textile wastewater. Rev. Chem. Eng. 2017, 33, 263–292. [Google Scholar] [CrossRef]
  103. Bao, J.; Yu, W.-J.; Liu, Y.; Wang, X.; Liu, Z.-Q.; Duan, Y.-F. Removal of perfluoroalkanesulfonic acids (PFSAs) from synthetic and natural groundwater by electrocoagulation. Chemosphere 2020, 248, 125951. [Google Scholar] [CrossRef]
  104. Liu, Y.; Hu, X.-M.; Zhao, Y.; Wang, J.; Lu, M.-X.; Peng, F.-H.; Bao, J. Removal of perfluorooctanoic acid in simulated and natural waters with different electrode materials by electrocoagulation. Chemosphere 2018, 201, 303–309. [Google Scholar] [CrossRef]
  105. Kamaraj, R.; Ganesan, P.; Lakshmi, J.; Vasudevan, S. Removal of copper from water by electrocoagulation process—Effect of alternating current (AC) and direct current (DC). Environ. Sci. Pollut. Res. 2013, 20, 399–412. [Google Scholar] [CrossRef]
  106. Yılmaz, A.E.; Boncukcuoğlu, R.; Kocaker, M.M.; Kocadağistan, E. An empirical model for kinetics of boron removal from boroncontaining wastewaters by the electrocoagulation method in a batch reactor. Desalination 2008, 230, 288–297. [Google Scholar] [CrossRef]
  107. Xi, Y.; Luo, Y.; Luo, J.; Luo, X. Removal of cadmium (II) from wastewater using novel cadmium ion-imprinted polymers. J. Chem. Eng. Data 2015, 60, 3253–3261. [Google Scholar] [CrossRef]
  108. Vasudevan, S.; Lakshmi, J.; Packiyam, M. Electrocoagulation studies on removal of cadmium using magnesium electrode. J. Appl. Electrochem. 2010, 40, 2023–2032. [Google Scholar] [CrossRef]
  109. Vasudevan, S.; Lakshmi, J.; Sozhan, G. Optimization of electrocoagulation process for the simultaneous removal of mercury, lead, and nickel from contaminated water. Environ. Sci. Pollut. Res. 2012, 19, 2734–2744. [Google Scholar] [CrossRef] [PubMed]
  110. Chithra, K.; Balasubramanian, N. Modeling electrocoagulation through adsorption kinetics. J. Model. Simul. Syst. 2010, 1, 124–130. [Google Scholar]
  111. Shafaei, A.; Rezaie, M.; Nikazar, M. Evaluation of Mn2+ and Co2+ removal by electrocoagulation: A case study. Chem. Eng. Process. Process Intensif. 2011, 50, 1115–1121. [Google Scholar] [CrossRef]
  112. Kabdaşlı, I.; Arslan, T.; Ölmez-Hancı, T.; Arslan-Alaton, I.; Tünay, O. Complexing agent and heavy metal removals from metal plating effluent by electrocoagulation with stainless steel electrodes. J. Hazard. Mater. 2009, 165, 838–845. [Google Scholar] [CrossRef]
  113. Kobya, M.; Demirbas, E.; Ulu, F. Evaluation of operating parameters with respect to charge loading on the removal efficiency of arsenic from potable water by electrocoagulation. J. Environ. Chem. Eng. 2016, 4, 1484–1494. [Google Scholar] [CrossRef]
  114. Rehman, A.; Kimb, M.; Reverberic, A.; Fabianoa, B. Operational parameter influence on heavy metal removal from metal plating wastewater by electrocoagulation process. Transactions 2015, 43, 2251–2256. [Google Scholar]
  115. Chou, W.-L.; Huang, Y.-H. Electrochemical removal of indium ions from aqueous solution using iron electrodes. J. Hazard. Mater. 2009, 172, 46–53. [Google Scholar] [CrossRef]
  116. Hashim, K.S.; Shaw, A.; Al Khaddar, R.; Pedrola, M.O.; Phipps, D. Iron removal, energy consumption and operating cost of electrocoagulation of drinking water using a new flow column reactor. J. Environ. Manag. 2017, 189, 98–108. [Google Scholar] [CrossRef]
  117. Moussa, D.T.; El-Naas, M.H.; Nasser, M.; Al-Marri, M.J. A comprehensive review of electrocoagulation for water treatment: Potentials and challenges. J. Environ. Manag. 2017, 186, 24–41. [Google Scholar] [CrossRef]
  118. Kobya, M.; Demirbas, E.; Bayramoglu, M.; Sensoy, M. Optimization of electrocoagulation process for the treatment of metal cutting wastewaters with response surface methodology. Water Air Soil Pollut. 2011, 215, 399–410. [Google Scholar] [CrossRef]
  119. Shaker, O.A.; Safwat, S.M.; Matta, M.E. Nickel removal from wastewater using electrocoagulation process with zinc electrodes under various operating conditions: Performance investigation, mechanism exploration, and cost analysis. Environ. Sci. Pollut. Res. 2022, 1–13. [Google Scholar] [CrossRef]
  120. Mateen, Q.S.; Khan, S.U.; Islam, D.T.; Khan, N.A.; Farooqi, I.H. Copper (II) removal in a column reactor using electrocoagulation: Parametric optimization by response surface methodology using central composite design. Water Environ. Res. 2020, 92, 1350–1362. [Google Scholar] [CrossRef]
  121. Khan, S.U.; Farooqi, I.H.; Usman, M.; Basheer, F. Energy efficient rapid removal of arsenic in an electrocoagulation reactor with hybrid Fe/Al electrodes: Process optimization using CCD and kinetic modeling. Water 2020, 12, 2876. [Google Scholar] [CrossRef]
  122. Al Aji, B.; Yavuz, Y.; Koparal, A.S. Electrocoagulation of heavy metals containing model wastewater using monopolar iron electrodes. Sep. Purif. Technol. 2012, 86, 248–254. [Google Scholar] [CrossRef]
  123. Thakur, L.S.; Goyal, H.; Mondal, P. Simultaneous removal of arsenic and fluoride from synthetic solution through continuous electrocoagulation: Operating cost and sludge utilization. J. Environ. Chem. Eng. 2019, 7, 102829. [Google Scholar] [CrossRef]
  124. Yilmaz, A.E.; Bayar, S.; Boncukcuoglu, R.; Fil, B.A. Removal of cadmium by electrocoagulation and a cost evaluation. Ekoloji 2012, 21, 26–33. [Google Scholar] [CrossRef]
  125. Nouri, J.; Mahvi, A.; Bazrafshan, E. Application of electrocoagulation process in removal of zinc and copper from aqueous solutions by aluminum electrodes. Int. J. Environ. Res. 2010, 4, 201–208. [Google Scholar]
  126. Ozyonar, F.; Karagozoglu, B. Operating cost analysis and treatment of domestic wastewater by electrocoagulation using aluminum electrodes. Pol. J. Environ. Stud. 2011, 20, 173. [Google Scholar]
  127. 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. 2015, 22, 127–131. [Google Scholar] [CrossRef]
  128. Dalvand, A.; Gholami, M.; Joneidi, A.; Mahmoodi, N.M. Dye removal, energy consumption and operating cost of electrocoagulation of textile wastewater as a clean process. Clean–Soil Air Water 2011, 39, 665–672. [Google Scholar] [CrossRef]
Figure 1. A schematic illustration of a conventional EC reactor.
Figure 1. A schematic illustration of a conventional EC reactor.
Sustainability 15 01708 g001
Figure 2. Schematic view of EC system with monopolar electrodes in parallel and monopolar electrodes in series and bipolar electrodes in series connections.
Figure 2. Schematic view of EC system with monopolar electrodes in parallel and monopolar electrodes in series and bipolar electrodes in series connections.
Sustainability 15 01708 g002
Table 1. Worldwide reports of some important trace metals, along with their source of contamination.
Table 1. Worldwide reports of some important trace metals, along with their source of contamination.
S. NoHeavy MetalsCountry/RegionSource TypeContamination Source/ReasonReference
1.Chromium, lead, copper, manganese and ironNorth GreeceSurface water used for drinkingEffluents discharged from industries and leaching with soil[11]
2.Arsenic, iron, lead, copper and zincAmiata and Siena, ItalyDrinking water suppliesLeaching with water-distributing pipes[12]
3.Chromium, lead, copper, cadmium and zincDipsiz stream, TurkeySurface waterCoal-fired power plant[13]
4.Arsenic and manganeseSouth Vietnam and CambodiaGroundwaterGeogenic[14]
5.Cadmium, chromium, lead, copper and nickelGreeceDrinking water suppliesGeogenic and anthropogenic activities[15]
6.Lead, arsenic, cadmium and zincDakahliya Governorate, EgyptTap waterCorrosion in plumbing fixtures[16]
7.ArsenicCoastal areas and 64 districts, BangladeshTube wells of shallow depthSubsurface contamination[17]
8.Chromium, lead, copper, cadmium and nickelBangladeshBuriganga riverEffluents from tanneries, other industrial outlets and pesticides[18]
9.Arsenic, chromium, and seleniumMakkah, Kingdom Saudi ArabiaGroundwaterGeological activity[19]
10.Cadmium and leadEgyptWater distribution systemBattery, steel and plastic making industries[20]
11.LeadSouthwest BangladeshRainwater harvesting structuresDissolutions from atmosphere[21]
12.Cadmium, LeadLagos State, Nigeriasurface waterscrap yard waste[22]
13.Cadmium, lead and ironKumasi district, GhanaGroundwater from boreholesIndustrial waste from paint industries, municipal sewage and anthropogenic waste[23]
14.Chromium, zinc, copper, manganese, nickel, lead and ironGoiânia, BrazilLeachateSanitary landfill[24]
15.Chromium, zinc, copper, manganese, nickel and ironNorth Mathiatis mine Nicosia, CyprusOpen cast mineMine drainage[25]
16.Nickel and ironQazvin, IranThermal power plantBoiler, air preheater and washing[26]
17.CopperEco-city, Bohai baySewage samplesWastewater treatment plant[27]
Table 2. Comparative assessment of heavy metal removal by various conventional methods adopted.
Table 2. Comparative assessment of heavy metal removal by various conventional methods adopted.
S. NoTreatment TechniqueTarget Metal IonEfficiency %AdvantagesDisadvantagesReferences
1.AdsorptionCd2+ & Cr6+55–60Low cost, simple techniqueModerately efficient, Regenerative power declines substantially[39]
2.Nano-adsorptionPb2+, Cr6+90–99%Selective removal, Highly efficientCostly, bulk production is not possible[40]
3.Coagulation and chemical precipitationCu2+, Cr3+, Pb2+ & Zn2+99.3–99.6Easily operational, Ease of sludge settling & dewateringCostly, high consumption of chemicals, huge sludge production[41]
4.Nano filtrationCu2+96Highly efficient, reliableLow anti-compacting ability[42]
5.Ion exchangeCu2+ & Ni2+99.14–99.33Selective removal, high regenerabilityCostly disposal problems with regenerative fluids[43]
6.ElectrochemicalNi2+, Cr6+98–100Highly efficient, no chemical requirementHigh operational cost[44]
Table 3. A sample of the studies that optimized operating conditions for heavy metal removal using the EC method.
Table 3. A sample of the studies that optimized operating conditions for heavy metal removal using the EC method.
S. NoMetal IonInitial Conc. (ppm)pHElectrode MaterialCurrent Density (A/m2) or Current (A)%Removal Efficiency References
1.Ni & Fe-8.1Fe-Fe1.5 A99[26]
2.Cr, Ni and Cu-6Al-Al1284.55, 89.65 and 95.16[27]
3.Cr and CuCu0 = 204 & 6Al-Al11.5793 and 99.4[53]
4.Fe, Zn, Mn, Cu, Ni, Cd and Cr-2.6Al-Al20 mA/cm2100[25]
5.SO42−-2.4Fe-Fe20010[54]
6.Cd-7Fe/Cu/Zn25 mA/cm299.73, 99.9, 82[55]
7.Cr8876Fe-Fe-100[56]
8.Ni3004–8Al-Al400100[57]
9.As(III)507.5Al-Al6092.2[58]
504Fe-Fe5.498.42[59]
10.Hg413–7Al-Fe4099.95[60]
203–7Al-SS3099[61]
11.Pb416Al-Fe2398.2[62]
27Fe-Fe8099.3[63]
12.Zn754–8Al-Al400100[64]
20.49.5Fe-Fe4097[52]
Table 4. Some used kinetic models to explain the adsorption of heavy metal ions in the EC process.
Table 4. Some used kinetic models to explain the adsorption of heavy metal ions in the EC process.
S. NoTarget Metal Ions (Pollutants)Anode EmployedKinetic ModelReference
1.Cu, Zn and NiIronPseudo-second-order[52]
2.Hg, Ni and PbMagnesiumSecond-order[109]
3.AsIron and AluminumFirst-order[110]
4.Co and MnAluminumFirst-order[111]
5.Ni and ZnStainless steelPseudo-first-order[112]
6.AsIronPseudo Second order[113]
7.Cu, Pb and NiAluminum and IronPseudo-first-order[114]
8.CdMagnesiumSecond-order[108]
9.PbMagnesiumSecond-order[63]
10.FeIronSecond-order[115]
Table 5. A summary of previous EC studies investigated the power consumption and operational cost with optimized conditions.
Table 5. A summary of previous EC studies investigated the power consumption and operational cost with optimized conditions.
S. NoMetalParametersRemoval (%)Energy ConsumptionOperating CostReferences
Local CurrencyUSD
1.CopperApplied current 0.26 A, pH 7, Time 5.4 min, Initial Concentration 27.8 ppm95%0.903 W-h/g Cu(II)0.134 INR/g removal0.002 US$/g[120]
2.ArsenicApplied current 0.46 A, pH 7, Initial concentration 10 ppm, Time 2 min95%3.1 W-h/g total As0.0974 INR/g removal0.0013 US$/g[121]
3.ChromiumApplied Current 1.48 A, pH 3, Time 21.47 min
Initial Concentration 49.96 ppm,
100%12.97 W-h/g
Cr (VI)
0.956 INR/g removal0.013 USD/g[28]
4.Copper, Zinc, Nickel and ManganeseCurrent density 25 mA/cm2, Time 50 min 49 kWh/m3€3.43/m33.77 USD/m3[122]
5.IronInitial concentration 20 ppm, Time 20 min, pH 6,
Current density 1.5 mA/cm2
98%3.2 kWh/m3 0.22 USD/m3[116]
6.Arsenic and FluorideRun time -95 min
Flow Rate -0.88 L/h
0.358 USD/m3[123]
7.Copper
Chromium
Nickel
pH 3.0, Time 20 min, Current 10 mA/cm2Cu-100%
Cr-100%
Ni-100%
10.07 kWh/m3 [44]
8.CadmiumInitial concentration 250 ppm, pH 6,
Current density 0.25 mA/cm2
58%0.363
kWh/m3
0.42 €/m30.46 USD/m3[124]
9.Zinc and CopperInitial Concentration 500 ppm, pH 7, Electric potential 40 V.Zinc-99.5%
Copper-99.7%
Zn- 0.095 kWh/g
Cu- 0.29 kWh/g
[125]
10.Domestic WastewaterCurrent density-100 A/m2, pH 7,
Time 10 min
COD- 72%
Turbidity-98%
Phosphorus- 98%
8.9 kWh/m3 0.86 USD/m3[126]
11.FluoridepH-7,
Current density
18.51 A/m2
85%3.43 kWh/m3 [127]
12.Textile WaterInitial concentration 50 ppm, electrode distance 1 cm, time 30 min98.59%1.303 kWh/m 0.256 USD/m3[128]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Khan, S.U.; Khalid, M.; Hashim, K.; Jamadi, M.H.; Mousazadeh, M.; Basheer, F.; Farooqi, I.H. Efficacy of Electrocoagulation Treatment for the Abatement of Heavy Metals: An Overview of Critical Processing Factors, Kinetic Models and Cost Analysis. Sustainability 2023, 15, 1708. https://doi.org/10.3390/su15021708

AMA Style

Khan SU, Khalid M, Hashim K, Jamadi MH, Mousazadeh M, Basheer F, Farooqi IH. Efficacy of Electrocoagulation Treatment for the Abatement of Heavy Metals: An Overview of Critical Processing Factors, Kinetic Models and Cost Analysis. Sustainability. 2023; 15(2):1708. https://doi.org/10.3390/su15021708

Chicago/Turabian Style

Khan, Saif Ullah, Mohammad Khalid, Khalid Hashim, Mehdi Hassanvand Jamadi, Milad Mousazadeh, Farrukh Basheer, and Izharul Haq Farooqi. 2023. "Efficacy of Electrocoagulation Treatment for the Abatement of Heavy Metals: An Overview of Critical Processing Factors, Kinetic Models and Cost Analysis" Sustainability 15, no. 2: 1708. https://doi.org/10.3390/su15021708

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop