You are on page 1of 25

Water Research 187 (2020) 116433

Contents lists available at ScienceDirect

Water Research
journal homepage: www.elsevier.com/locate/watres

Review

Electrode passivation, faradaic efficiency, and performance


enhancement strategies in electrocoagulation—a review
Markus Ingelsson a, Nael Yasri a, Edward P.L. Roberts a,∗
a
Department of Chemical & Petroleum Engineering, University of Calgary, 2500 University Dr NW, Calgary, AB T2N 1N4, Canada

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

Article history: Treating water and wastewater is energy-intensive, and traditional methods that require large amounts of
Received 17 June 2020 chemicals are often still used. Electrocoagulation (EC), an electrochemical treatment technology, has been
Revised 30 August 2020
proposed as a more economically and environmentally sustainable alternative. In EC, sacrificial metal
Accepted 15 September 2020
electrodes are used to produce coagulant in-situ, which offers many benefits over conventional chem-
Available online 16 September 2020
ical coagulation. However, material precipitation on the electrodes during long term operation induces
Keywords: a passivating effect that decreases treatment performance and increases power requirements. Overcom-
Electrocoagulation ing this problem is considered to be the greatest challenge facing the development of EC. In this critical
Water treatment review, the studies that have examined the nature of electrode passivation, and its effect on treatment
Electrochemistry performance are considered. A fundamental approach is used to examine the association between passi-
Passivation vation and faradaic efficiency, a surrogate for EC performance. In addition, the strategies that have been
Faradaic efficiency
proposed to remove or avoid passivation are reviewed, including aggressive ion addition, AC current oper-
Polarity reversal
ation, polarity reversal, ultrasonication, and mechanical cleaning of the electrodes. It is concluded that the
success of implementing each method is dependent on critical operating parameters, and careful consid-
eration should be taken when designing an EC system based on the phenomena discussed in this article.
In conclusion, this review provides insight into passivation mechanisms, delivers guidelines for sustaining
high treatment performance, and offers an outlook for the future development of EC.
© 2020 Elsevier Ltd. All rights reserved.

1. Introduction from the cathode (Yasri et al., 2020). The cationic character of
the metal-oxyhydr(oxides) neutralizes the negative surface charge
Electrocoagulation (EC) is a water and wastewater treatment of colloidal contaminants, which causes them to destabilize and
technology that produces metal coagulant in-situ by electro- form larger aggregates—together with the metal-oxy(hydroxides)—
chemically dissolving sacrificial electrodes that are submerged that can be separated by filtration, flotation or sedimentation
in contaminated water. This is achieved by applying a current (Garcia-Segura et al., 2017).
to the electrodes from an external power source (Vik et al.,
1984). The positively polarized anode, commonly made of steel M(s ) → Mz+ (aq) + ze− (1)
(Fe-EC) or aluminum (Al-EC), is electrochemically dissolved Eq.
(1) (Yasri and Gunaskearan, 2017). At the same time, H2 O is 2H2 O + 2e− → 2OH− (aq) + H2(g ) (2)
electrolyzed at the negatively charged cathode to produce OH−
and H2(g) Eq. (2) (Mansoorian et al., 2014). Following anodic
dissolution, metal cations hydrolyze to form oxy(hydr)oxides and Mz+ (aq) + zH2 O → M(OH )z(s ) + zH+ (aq) (3)
H3 O+ ions (symbolized as H+ in this article) that cause the pH
where M represents the metal of which the electrode is made
to drop Eq. (3) (Lakshmanan et al., 2009). The acidic and alka-
of, and z is the charge transfer number. An illustration of the EC
line environments at the electrodes create distinctive layers that
process is shown in Fig. 1.
have shown to be up to 0.5 mm thick using confocal imaging
The amount of electrochemically produced coagulant in EC
(Fuladpanjeh-Hojaghan et al., 2019). Subsequently, H+ diffuses
can be predicted by Faraday’s Law of Electrolysis (Eq. (4)) and
to the electrolyte, where it is neutralized by OH− transported
is proportional to the amount of charge passed through the cell
(Yasri and Gunaskearan, 2017). The majority of EC experiments

Corresponding author. reported in the literature are operated galvanostatically—i.e., at
E-mail address: edward.roberts@ucalgary.ca (E.P.L. Roberts). constant current—to ensure effective addition of the coagulant.

https://doi.org/10.1016/j.watres.2020.116433
0043-1354/© 2020 Elsevier Ltd. All rights reserved.
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

izing chemical such as lime, Ca2 (OH)2 (Garcia-Segura et al., 2017).


Nomenclature The cathodic reaction in EC produces H2(g) and OH− , providing an
in-situ pH neutralizing mechanism (Wang et al., 2009). Further-
EC electrocoagulation more, the uniform metal dissolution of the electrodes (as opposed
CC chemical coagulation to point-addition in CC) reduces the need for intense mechani-
Al-EC electrocoagulation with aluminum electrodes cal mixing, which lowers energy demands while making the co-
Fe-EC electrocoagulation with iron electrodes agulation process more efficient. Uniform metal ion addition in EC
AC-EC alternating current electrocoagulation also permits lower coagulant dosing, which further curtails sludge
DC-EC direct current electrocoagulation production. Additionally, the prospect of harnessing the H2(g) pro-
PR-EC polarity reversal electrocoagulation duced at the cathode makes EC a viable energy storage technology
PRT polarity reversal time (Phalakornkule et al., 2010). Lastly, using commonplace materials
Sono-EC ultrasonication electrocoagulation such as iron and aluminum for local electrode manufacturing al-
lows remote communities to sustain treatment without relying on
resupplying chemicals, thus eliminating transportation costs that
The faradaic efficiency is often reported when evaluating EC per- are burdensome to areas such as Northern Canada (Ragush et al.,
formance, which is the ratio of experimentally observed coagu- 2015; Tartakovsky et al., 2018).
lant mass over the theoretical mass predicted by Faraday’s Law However, problems with long-term EC operation persist. For ex-
(Eq. (5)). The faradaic efficiency is determined by measuring the ample, the precipitation of solids on the electrodes over time—and
amount of metal released over the duration of the experiment. This its ensuing passivating effect—increases operating costs and low-
is generally performed by dissolving all flocs in acid, followed by ers treatment efficiency (Bandaru et al., 2020; Garcia-Segura et al.,
analysis using inductively coupled plasma optical emission spec- 2017; Müller et al., 2019). Furthermore, storing the electrodes
troscopy (ICP-OES) (Chow and Pham, 2019; Dubrawski et al., 2015). between treatment cycles and removing the initial passivating
A calculated faradaic efficiency of 100% is desirable since it is in- surface layer during reactor startup has shown to be problem-
dicative of effective current utilization. atic (van Genuchten et al., 2016). Additionally, non-uniform elec-
ItM trode consumption can compromise the structural integrity of
Theoretical amount of coagulant metal dissolved = (4) the electrodes, which requires them to be prematurely exchanged
zF
(Bandaru et al., 2020; Wellner et al., 2018). It is widely recognized
Observed Coagulant zF m that overcoming these challenges is imperative to the development
FE[%] = = × 100% (5)
Theoretical Coagulant ItM of EC (Bian et al., 2019; Kabdaşlı et al., 2012).
In this article, the existing body of scientific literature related
where F is Faraday’s constant (96,485 C mol−1 ), m is the exper-
to the nature of electrode passivation in EC is reviewed. A funda-
imentally observed mass (g), I is the current intensity (A), t is
mental approach is used to study the electrochemical and chemi-
the electrolysis time (s), and M is the molar mass of the metal (g
cal mechanisms governing coagulant production and electrode pas-
mol−1 ).
sivation, while also examining different strategies that have been
EC is often compared to chemical coagulation (CC), where metal
proposed to mitigate the detrimental effects of passivation. Some
ions are added to contaminated water as salts such as aluminum
of these strategies include increasing the flowrate to enhance hy-
sulfate or ferric chloride. However, it has been shown that EC is
drodynamic scouring (Karamati-Niaragh et al., 2019; Timmes et al.,
more cost-effective than CC per unit water treated for many differ-
2010), adding depassivating salts to the water (van Genuchten
ent applications because of some significant advantages that EC ex-
et al., 2016), mechanically cleaning the electrodes (Sahu et al.,
hibits (Garcia-Segura et al., 2017; Price et al., 2018; Timmes et al.,
2014), ultra-sonication (Maha Lakshmi and Sivashanmugam, 2013)
2010; Vik et al., 1984). For example, the cationic coagulant in
or changing the current waveform (Eyvaz et al., 2009; Karamati-
EC is produced in-situ without adding unnecessary counter-ions
Niaragh et al., 2019; Pi et al., 2014; Sahu et al., 2014; Timmes et al.,
that contribute to sludge production (Ölmez, 2009). Meanwhile,
2010; van Genuchten et al., 2017, 2016; Yet, 2011). Each method is
counter-ions in CC remaining in the dissolved phase must be re-
evaluated, and suggestions for future research are provided.
moved in subsequent treatment steps depending on discharge lim-
To date, published reviews of EC have focused on general op-
its (Mamelkina et al., 2017). In addition, neutralizing chemicals
eration or on the abatement of a specific pollutant (Chen, 2004;
must be added in CC to mitigate the pH drop caused by the hydrol-
Emamjomeh and Sivakumar, 2009; Garcia-Segura et al., 2017;
ysis of the metal salts (Eq. (3)). This is achieved by using a neutral-
Kabdaşlı et al., 2012; Mollah et al., 2001, 2004a; Moussa et al.,
2017; Sahu et al., 2014; Yasri et al., 2020). While these reviews are
useful, pollutant removal is dependent on both coagulant dosing
in the EC reactor, and the subsequent separation process. Overall,
this involves many controllable variables, which makes it difficult
to distinguish between the sensitivity of different operating param-
eters. As concluded by Timmes et al., 2010, EC should be regarded
as a two-step process, i.e., coagulant dosing followed by floccula-
tion and separation. Therefore, the scope of this article is to fo-
cus on the coagulant dosing component, while disregarding floc-
separation.

2. Electrode surface layers, passivation, and EC performance

2.1. Principles of electrode surface layer formation and passivation

Passivation is a concept emanating from corrosion science,


which describes the formation of a protective film on a metal
Fig. 1. Schematic diagram of the electrocoagulation (EC) process. surface, usually comprised of metal oxides. The passivation layer

2
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Fig. 2. The distribution of the cell potential in the EC reactor.

impedes the kinetics of otherwise thermodynamically favorable tion (Huang et al., 2009). For example, anodic O2(g) or Cl2(g) evo-
metal oxidation, which increases the resistance at the electrode- lution (Eqs. (7) & (8)) have been proposed as unwanted side reac-
electrolyte interface (Schmuki, 2002). In the EC literature, this is tions that occur when the total anodic overpotential is sufficiently
commonly referred to as a surface layer film (Müller et al., 2019; high during galvanostatic operation (Dubrawski et al., 2015).
van Genuchten et al., 2016). The terms “passive layer” and “sur-
face layer” are often used interchangeably. However, a surface layer 2H2 O ↔ O2 + 4H+ + 4e− E o = 1.23 V vs SHE (7)
does not have to be passivating, i.e., it could also be conductive
(Bandaru et al., 2020; Schmuki, 2002). Furthermore, the surface 2Cl− ↔ Cl2 + 2e− E o = 1.36 V vs SHE (8)
layer formation process has also been referred to as electrode foul-
ing (Schulz et al., 2009; Yasri et al., 2020). The formation of the electrode surface layer is dependent on
To sustain effective electrochemical dissolution of the anode, specific EC operating conditions such as electrolyte conductiv-
the applied cell potential (U) must overcome several resistances ity/composition, electrode material, dissolved oxygen (DO), current
within the cell (Eq. (6)). These energy barriers are expressed density, and polarization cycles (van Genuchten et al., 2017, 2016).
as a series of overpotentials (η), which are encountered at the During EC operation, it is generally desirable to maintain a high
electrode-electrolyte interfaces—namely, the activation, concentra- electrolyte conductivity and a low inter-electrode distance to re-
tion, and passivation overpotentials (Chen et al., 2002). The acti- duce ohmic losses (potential drop) in the solution, which reduces
vation overpotential is associated with the kinetic energy barrier the cell potential (Eq. (6)) (Mollah et al., 2001). However, small
that the electron-exchange process exhibits, whereas the passiva- inter-electrode gaps and high electrolyte conductivities increase
tion overpotential pertains to the energy required for dissolved the variations in the current density distribution over the electrode
metal to penetrate the passivating surface layer, which increases geometry, which causes uneven wear of the anode and is problem-
the polarization resistance. Due to the availability of metal in the atic to long-term operation (McBeath et al., 2020).
anode, the concentration overpotential there is negligible. The electrode surface layer can be a mono-layer of absorbed
molecules, or consist of multiple layers (Schmuki, 2002). For ex-
d
U = E + ηan,a + ηan,c + ηan,p + |ηct,a |+|ηct,c | + |ηct,p | + i (6) ample, a porous surface layer will be less passivating than a non-
k porous layer due to its enhanced charge carrying properties. Sev-
where E is the thermodynamic equilibrium potential for the elec- eral fouling mechanisms are possible and may include the pre-
trochemical cell, and ηan,a , ηan,c , and ηan,p are the anodic acti- cipitation of dissolved species, adsorption of colloidal/suspended
vation, concentration, and passivation overpotentials, respectively. particles, or solid-phase transformations at the anode surface
Likewise, ηct,a , ηct,c , and ηct,p are the equivalent overpotentials at (Pally et al., 2020). The formation of a metal oxide layer through
the cathode. Furthermore, d is the distance between the electrodes, anodization to prevent further corrosion is an established phe-
k is the electrolyte conductivity, and i is the current density. Fig. 2 nomenon for aluminum, zinc, magnesium, and several other met-
illustrates the distribution of the cell potential and breaks down als (Roberge, 2008). For example, metal oxide coated anodes are
the components outlined in Eq. (6) within the EC reactor. The equi- utilized for some electrochemical applications, e.g., anodic electro-
librium potential depicted at the anode is relative to the potential oxidation for wastewater disinfection (Wu et al., 2014).
of the cathode. At the anode interface, the low pH environment inhibits the
As an anodic surface layer is formed, metal dissolution is buildup of a surface layer, which works as a passivation prevention
impeded, which raises the passivation overpotential and in- mechanism. However, metal oxides can precipitate on the anode
creases the energy consumption in EC Holt et al., 2005, 2002; over time if their solubility limits are exceeded (Chen et al., 2002).
Mechelhoff et al., 2013a; van Genuchten et al., 2016; Xu et al., Similarly, Ca2+ and Mg2+ in the electrolyte can precipitate as min-
2015; Yang et al., 2015). Extreme passivation may lead to elec- erals and form a passivating surface layer in the high pH environ-
trochemical side reactions that decrease coagulant production dur- ment at the cathode (Doggaz et al., 2019; Sahu et al., 2014). Addi-
ing galvanostatic operation or cause a complete termination of the tionally, cations could be transported and sorbed to the negatively
passage of current during potentiostatic (constant voltage) opera- polarized cathode by electrophoretic migration and electrostatic

3
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Fig. 3. Fouling processes at the anode and the cathode.

interactions (Sahu et al., 2014; Xu et al., 2017), otherwise referred


to as electro-condensation (Hu et al., 2003). Likewise, anions could
migrate towards—and bind to—the positively polarized anode. This
effect could explain why electrode surface layers have been ob-
served to be thicker at higher current densities (Dubrawski et al.,
2015; Mansouri et al., 2011; Müller et al., 2019), although the effect
of electrophoretic migration is diminished at high supporting elec-
trolyte concentrations (Mechelhoff, 2008). Alternatively, electro-
chemically produced metal cations could precipitate with oxyan-
ions and adsorb to the electrode surface. For example, a passivat-
ing AlPO4 surface layer has been observed on electrodes in phos-
phate electrolytes in Al-EC (Mansouri et al., 2011). Similarly, FeSO4
can form a passivating surface layer in the presence of SO4 2− in
Fe-EC (Gerónimo-López et al., 2014). The probability of exceed-
ing the solubility limits of metal salts at the anode interface is
greater while operating at higher current densities because of the
increased concentration of metal ions. The exact mechanisms for
how co-existing ions contribute to passivation in EC is not clear,
but an illustration of the possible mechanisms is provided in Fig. 3.
In batch experiments, it is often seen that the bulk solution Fig. 4. Tafel representation of potentiodynamic polarization data, a useful tool for
pH shifts towards higher values over time, especially at higher understanding the corrosion characteristics of metal in a specific electrolyte. Exper-
imental conditions: aluminum working electrode, platinum counter electrode, SCE
current densities due to OH− production outcompeting metal- reference electrode, pH 9, 0.1 M NaCl, 0.5 mV s−1 , T = 25 °C. Reprinted (adapted)
hydrolysis—i.e., H+ generation—which has more sluggish reaction with permission from Mansouri et al., 2011. Anodic dissolution of pure aluminum
kinetics (Garcia-Segura et al., 2017; Lakshmanan et al., 2009). during electrocoagulation process: Influence of supporting electrolyte, initial pH,
For example, the pH equilibrated around 8–9 in Al-EC after 60– and current density. Ind. Eng. Chem. Res. 50, 13,362–13,372. Copyright (2011) Amer-
ican Chemical Society.
80 min of galvanostatic operation, which was independent of the
initial pH (Mansouri et al., 2011). Additionally, when buffering
agents are present—i.e., salts of weak acids such as phosphates
(pKa1 = 2.1, pKa2 = 7.2, pKa1 = 12.4) and carbonates (pKa1 = 3.6, ther be beneficial or detrimental to avoiding passivation, depend-
pKa2 = 6.3, pKa3 = 10.3)—H+ protons produced through hydroly- ing on the solution composition.
sis at the anode are quenched. Consequently, the acidic pH bound-
ary layer is thinner in buffered solutions than in unbuffered sys- 2.2. Potentiodynamic polarization methods to study passivation
tems, which increases the likelihood of metal-oxyhydr(oxides) pre-
cipitating at the anode (Lumsden et al., 1981; Refaey, 2005). The nature of electrode passivation can be characterized by po-
However, it has been shown that adding salts containing NH4 + larization studies (Dura and Breslin, 2019; Schmuki, 2002). Poten-
(pKa = 9.3), which is amphoteric, enhances EC performance in tiodynamic polarization methods are often used in corrosion stud-
Al-EC by avoiding an excessive increase in bulk solution pH ies (Lee and Pyun, 20 0 0; Pyun et al., 1999). However, some studies
(Izquierdo et al., 2010; Trompette and Vergnes, 2009). It has also have adopted this methodology to investigate the characteristics of
been shown that the shift towards alkaline pH is less pronounced electrochemical dissolution in EC (Hu et al., 2003; Mansouri et al.,
in Al-EC than in Fe-EC due to the buffering effect of aluminum 2011; Panikulam et al., 2018). Data from these experiments are
monomeric species that form in the bulk solution ([Al(OH)]2+ , usually presented as Tafel plots (Fig. 4). The minimum in current
[Al(OH)2 ]+ , [Al2 (OH)2 ]4+ , [Al(OH)4 ]− ), and polymeric species such density determines the equilibrium potential (E), also referred to
as [Al6 (OH)15 ]3+ , [Al7 (OH)17 ]4+ , [Al8 (OH)20 ]4+ , [Al13 O4 (OH)24 ]7+ as the corrosion potential (Esmailzadeh et al., 2018). At this poten-
(Daneshvar et al., 2007; Fayad et al., 2017; Kobya et al., 2006). Ul- tial, the kinetic rates of reduction/oxidation are equal, and the net
timately, the presence of buffering agents in the electrolyte can ei- current approaches zero. The current density plateau in the anodic

4
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

potential region—adjacent to the corrosion potential—is the passi-


vation range. Here, the presence of a passivating surface layer im-
pedes the dissolution of the electrode, even though electrochemi-
cal oxidation becomes more thermodynamically favorable with in-
creasing anodic potentials. However, the current dramatically in-
creases at a certain potential, indicating the start of the active dis-
solution range, also referred to as pitting corrosion (Zhang and
Ma, 2019). The width of this range corresponds to the passivation
overpotential, and an increase in the passivation range is indica-
tive of a higher polarization resistance at the electrode-electrolyte
interface.
Pitting is initiated by ions breaking down the electrode surface
layer, or by the dissolution of the surface film when the electrode
boundary layer becomes sufficiently acidic due to metal hydrolysis
(Lumsden et al., 1981). Aggressive ions such as Cl− enhance pitting
corrosion due to the ionic mobility that allows them to penetrate
the electrode surface oxide film. Consequently, Cl− reacts with the
metal oxides, which dissolves the surface layer and exposes the
underlying raw metal (Golder et al., 2007). It has also been pro-
posed that ClO− (formed at the anode following Cl− →Cl2 oxida- Fig. 5. Pourbaix diagram of iron in water at 25 °C; the dashed lines represent
the electrochemical stability range of water. Produced based on calculations by the
tion) destroys the passivating surface layer (Zhou et al., 2020). In MEDUSA-HYDRA software (Puigdomenech, 2019).
the presence of Cl− , the corrosion and pitting potentials shift to-
wards more negative potentials while decreasing the width of the
passivation range (Mansouri et al., 2011). 2.3. Passivation of different electrode materials
It is well established that the electrode corrosion character-
istics are dependent on the nature of the supporting electrolyte The mechanism and rate of electrode passivation and corro-
(Hu et al., 2003; Mansouri et al., 2011). By changing operating sion vary depending on the electrode materials that are used.
conditions such as the electrolyte composition, pH, or electrode EC is traditionally operated with either steel or aluminum elec-
material—the corrosion and pitting potentials shift. Higher corro- trodes. Other metals such as zinc (Liu et al., 2018; Vasudevan and
sion/pitting potentials are indicative of increased passivation, i.e., Lakshmi, 2012), magnesium (Kamaraj et al., 2013) and titanium
a greater voltage must be applied to initiate metal dissolution. (Ti) (El-Ghenymy et al., 2020) have also been shown to abate
Additionally, based on operating conditions, the current densities aqueous contaminants effectively using EC, but they are also sus-
within a potential window vary. For example, current densities ceptible to passivation (El-Ghenymy et al., 2020; Kamaraj et al.,
within the scanned potential window were higher in Fe-EC than 2013; Vasudevan and Lakshmi, 2012). Most research has focused
in Al-EC in a synthetic kaolin solution, indicating faster corrosion on studying the electrochemical behavior, speciation, and coag-
kinetics (Panikulam et al., 2018). Comparatively lower current den- ulation mechanisms for iron (usually steel) and aluminum elec-
sities in the active dissolution range are also indicative of greater trodes due to their abundant availability and low procurement
polarization resistance and electrode passivation. costs (Holt et al., 2002). The price of steel is more economi-
Hu et al., 2003 reported greater polarization resistance with cal than aluminum, but better removal performance with alu-
aluminum electrodes when scanning from low to high potentials minum electrodes for certain pollutants has made it an attractive
(−1.5 V to 3.0 V [vs. SCE]) in the presence of only F− , compared alternative.
to when Cl− , NO3 − , and SO4 − were present in the electrolyte. It
was hypothesized that enhanced electrophoretic migration and ad- 2.3.1. Passivation behavior of iron electrodes in EC
sorption of F− to the anode caused this effect. In contrast, no pas- Observed rates of metal dissolution indicate that Fe(II) ions
sive region was observed under similar experimental conditions, are initially dissolved from the Fe(0) anode (Eq. (9)), which is
but without F− in the electrolyte (Mansouri et al., 2011). thermodynamically favorable over Fe(0)→Fe(III) oxidation (Fig. 5).
Corrosion characteristics are also affected by the pH at the Fe(II) is subsequently oxidized to Fe(III) ions by dissolved
electrode-electrolyte interface. For example, by increasing the ini- oxygen in the bulk solution (Eq. (10)), which is favored at
tial pH from 2 to 11 in a NaCl electrolyte (0.1 M), it was ob- higher pH (Dubrawski and Mohseni, 2013; Holt et al., 2002;
served that the corrosion potential shifted towards more negative Lakshmanan et al., 2009; Sasson et al., 2009; van Genuchten et al.,
potentials when scanning from −1.5 V to 3.0 V (vs. SCE) in Al- 2017; Vepsäläinen, 2012). Additionally, given a sufficiently acidic
EC (Mansouri et al., 2011). In the same study, Mansouri et al. ob- environment, Fe(0) can be chemically dissolved from the anode
served that the pitting potential was independent of pH, which (Eq. (11)) (Golder et al., 2011).
corresponded to the increased width of the passivation range as
the pH was increased. A similar observation was also made by El- Fe(s ) ↔ Fe2+ (aq ) + 2e− E o = −0.44 V vs SHE (9)
Ghenymy et al., 2020 with titanium electrodes. Both studies at-
tributed the changes in the corrosion characteristics to a more fa- 4Fe2+ + O2 + 2H2 O → 4Fe3+ + 4OH− (10)
vorable dissolution of the electrode in alkaline pH. However, a shift
towards more negative corrosion potentials (when scanning from
Fe0 + 2H+ → Fe2+ + H2 (11)
cathodic to anodic currents) is also caused by the cathodic elec-
trolysis of H2 O being more thermodynamically favorable in lower Following the generation of these ions, they hydrolyze into
pH (Roberge, 2008). This effect explains why the corrosion poten- a wide variety of oxyhydr(oxides)—which can be comprised of
tials were relatively high at pH 2, even though aluminum dissolu- Fe(II), Fe(III), or a combination of both (Dubrawski et al., 2015;
tion was favorable under these conditions (see Section 2.3.2). The Lakshmanan et al., 2009). Some oxyhydr(oxides), such as magnetite
corrosion parameters reported from potentiodynamic polarization (Fe3 O4 ) and hematite (Fe2 O3 ), can form by the solid-phase oxida-
studies in EC are presented in Table 1. tion of Fe(0) in alkaline pH. Furthermore, Fe-oxyhydr(oxides) func-

5
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Table 1
Reported values for corrosion parameters from potentiodynamic polarization investigations for different electrode materials and electrolyte compositions in EC.

Electrode
Material Sample Type Corrosion Potential (V vs SCE) Pitting Potential (V vs SCE) Current Density (i) Author

Fe 5 mM K2 Cr2 O7 , 0–19 mM 5 mM K2 Cr2 O7, 5 mM K2 Cr2 O7 + Passivation (Xu et al.,


NaCl, K2 SO4 , pH 8 0–19 mM NaCl: incurred when Cl− 2015)

• 18 & 19 mM Cl : −0.41 V
−1.1V conc was lower
• 17 mM Cl− : −0.21 V
than 16 mM
• 0 & 16 mM Cl− : 0.84V

Al 0.47 M F− with: Highest i for F− & (Hu et al.,


Cl− . Lowest i for DI 2003)
• 5 mM Cl − • DI H2 O: −0.8 V • DI H2 O: No pitting
H2 O.
• 5 mM NO3 − • F− & Cl− : −1.3 V • F− & Cl− : −0.5 V
• 5 mM SO4 2− • F− & SO4 2− : −1.3 V • F− & SO4 2− : No pitting
• F− & NO3 − : −1.0V • F− & NO3 − : 1.8 V

Al Lower i for Na2 SO4


& NaH2 PO4 . High i (Mansouri et al.,
• 0.1 M NaCl • DI H2 O: −0.668 V • DI H2 O: No pitting
for NaCl and DI 2011)
• 0.1 M Na2 SO4 • NaCl: −1.213 V • NaCl: −0.646 V
H2 O.
• 0.1 M NaH2 PO4 • Na2 SO4 : −1.056 V • Na2 SO4 : ~0.4 V
• in DI Water (pH 9) • NaH2 PO4 : −0.519 • NaH2 PO: No pitting

Al 0.5 mM Na2 SO4 : Higher i in NaCl


−1.45 V soln. No i increase (Mechelhoff et al.,
• 0.5 mM Na2 SO4 , • 0.5 mM Na2 SO4 , 0.4 mM
in soln with only 2013a)
• 0.4 mM NaCl, NaCl: −0.52 V Na2 SO4 . Humic
• 5–10 g L−1 humic acid • 0.5 mM Na2 SO4 , 0.4 mM acid lowers i
• in pH 7 RO H2 O NaCl, 5 g L−1 humic acid:
• −0.52 V
• 0.5 mM Na2 SO4 , 0.4 mM
NaCl, 10 g L−1 humic acid:
• −0.52 V
• 0.5 mM Na2 SO4 : No pitting

Al – Mg Little/no passivation region, i.e., Little change in (Dura and


pitting initiated at corrosion corrosion current Breslin, 2019)
• 0.170 M NaCl • High NaCl: −0.664 V
potential between high/low
• 0.017 M NaCl • Low NaCl: −0.554 V
NaCl conc. KH2 PO4
• 0.017 M • NaCl + KH2 PO4 : −0.5V
decreased
NaCl + 8.12 × 10−4 corrosion current
KH2 PO4 by 35%

Ti Little change in (El-


corrosion current Ghenymy et al.,
• 0.5 M Na2 SO4 • Na2 SO4 : −0.325 V • Na2 SO4 : ~0.75 V
between high/low 2020)
• 0.5 M NaCl • NaCl: −0.398 V • NaCl: ~0.7 V
NaCl conc
• 0.5 M Na2 HPO4 • Na2 HPO4 : −0.040 V • Na2 HPO4 : —
• 0.5 M (NaCl + Na2 SO4 ) • NaCl + Na2 SO4 : 0.405 V • NaCl + Na2 SO4 : ~0.2 V
• 0.5 M • NaCl + Na2 SO4 + Na2 HPO4 : • NaCl + Na2 SO4 + Na2 HPO4 :
(NaCl + Na2 SO4 + Na2 HPO4 ) −0.386V ~0V

tion as the coagulant in the bulk solution, but also contribute to trolyte composition, and the [Fe(II)]:[Fe(III)] ratio within the re-
the buildup of the electrode surface layer (van Genuchten et al., actor. The nature of these parameters varies based on the charge
2017). Therefore, the nature of Fe(II/III) speciation is vital because loading rate (C L−1 min−1 ), which is a function of the current
different oxyhydr(oxides) exhibit distinctive passivating character- density and reactor geometry. Because green rust and Fe3 O4 are
istics (Bandaru et al., 2020). The formation of different oxides is Fe(II/III) complexes, they form when the [Fe(II)]:[Fe(III)] ratio is
best represented by Pourbaix diagrams (Fig. 5), which show the low, i.e., in acidic and low DO conditions when Fe(II)→Fe(III) ox-
stability range of metal oxides with respect to electrochemical and idation is retarded (Dubrawski et al., 2015). Additionally, green
pH conditions (Schmuki, 2002). rust is an intermediate in Fe3 O4 and γ -FeOOH formation and
The speciation of Fe(II/III)-oxyhydr(oxides), depending on DO can be isolated in the presence of sulfates, phosphates, and car-
and electrolyte chemistry, has been mapped and is shown in Fig. 6 bonates (Dubrawski et al., 2015; Moreno et al., 2009). Further-
(Dubrawski et al., 2015). It has been demonstrated that the pri- more, Fe(II/III)-oxyhydr(oxide) formation is favored when the rate
mary Fe-oxyhydr(oxides) that form in EC are green rust, Fe3 O4 , and of Fe(II) generation is at least three times higher than the rate of
lepidocrocite (γ -FeOOH) (Dubrawski et al., 2015; Dubrawski and DO replenishment (van Genuchten et al., 2018). Therefore, if O2(g)
Mohseni, 2013). Of these products, Fe(III) species (γ -FeOOH) ex- is produced in anodic side reactions, DO levels would increase and
hibit the poorest conductivity, and their incorporation in the elec- prompt Fe(II)→Fe(III) oxidation, hence facilitating the formation of
trode surface layer is conducive to a higher degree of passivation Fe(III)-oxyhydr(oxides), i.e., γ -FeOOH (Fig. 6). By analyzing the an-
(Bandaru et al., 2020; Gerónimo-López et al., 2014). Meanwhile, ode surface layer using X-ray powder diffraction (XRD), it was de-
Fe(II)-containing species (e.g., Fe3 O4 ) are more conductive and less termined that Fe3 O4 is its main constituent (Bandaru et al., 2020;
detrimental to passivation (Bandaru et al., 2020). van Genuchten et al., 2016). Consequently, this makes the com-
It has also been reported that the formation of the differ- position of the electrode surface layers in EC different than the
ent precipitates depends on the concentration of DO, pH, elec- surface layers on naturally corroded steel, which usually contain

6
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Fig. 6. Diagram for the speciation of iron depending on operating conditions. Reprinted (adapted) with permission from Dubrawski et al., 2015. Production and transforma-
tion of mixed-valent nanoparticles generated by Fe(0) electrocoagulation. Environ. Sci. Technol. 49, 2171–2179. Copyright (2015) American Chemical Society.

with Fe(II/III) complexes (van Genuchten et al., 2017). However,


NO3 − accelerates Fe-oxyhydr(oxide) formation, which enhances the
buildup of the electrode surface layer. This effect may be caused
by the electrochemical reduction of NO3 − to NO2 − , NH3 , and N2 ,
which liberates hydroxide ions and raises the solution pH. Ad-
ditionally, in solutions with considerable sulfate concentrations,
Fe(II/III) ions have shown to form precipitates with SO4 2 − . For ex-
ample, FeSO4 and green rust sulfate may form on the electrode
surface, which contributes to passivation (Gerónimo-López et al.,
2014). The ordering of ionic species from least to most negative
impact on Fe(0) dissolution (and passivation) has been determined
as chloride < bromide < sulfate < formate < carbonate < phos-
phate < citrate < nitrate (van Genuchten et al., 2017).
The effect of increasing current density has also been shown
Fig. 7. The effect of varying electrolyte composition on the anodic potential in Fe- to affect passivation in Fe-EC. For example, high current densities
EC, 1 mA cm−2 , initial pH 7; concentrations: nitrate (10 mM), sulfate (5 mM), NaCl increase the total overpotential (η) at the electrode according to
(10 mM). Van Genuchten et al., 2017. Factors affecting the faradaic efficiency of
the Butler-Volmer equation (Eq. (12)) (Dubrawski et al., 2015).
Fe(0) electrocoagulation. J. Environ. Chem. Eng. 5, 4958–4968.
  a zF   a zF 
i = i0 exp
a
η − exp c η (12)
RT RT
Fe(III) species such as γ -FeOOH, akaganeit (b-FeOOH), or hematite where i is the current density, i0 is the exchange current density,
(a-Fe2 O3 ) (Cook, 2005; van Genuchten et al., 2016). aa and ac are electron transfer coefficients, R is the universal gas
Furthermore, it is well-known that little passivation occurs constant, T is the temperature, and η is the total electrode over-
in solutions with high Cl− concentrations. In contrast, electrodes potential (activation, concentration and passivation overpotentials).
are prone to passivate in the presence of dissolved oxyanions When the anode overpotential corresponding to the applied cur-
such as nitrate, sulfate, carbonate, phosphate, silicate, formate, rent density is sufficiently high, O2(g) evolution can be triggered.
citrate, molybdate, and chromate (Lakshmipathiraj et al., 2008; This side reaction has been seen when operating at 125 mA cm−2
Refaey, 2005; van Genuchten et al., 2017; Xu et al., 2015). As an (Dubrawski et al., 2015), which is exceptionally high compared
electrode surface passivates or depassivates, the anodic overpo- to the typical current densities reported by other authors (see
tential increases or decreases, respectively. This response becomes Tables 2-4). Additionally, a higher current density leads to more
apparent in chronopotentiometric and chronogalvanometric stud- significant electrophoretic migration of charged suspended parti-
ies (Fig. 7) (Mechelhoff et al., 2013a; van Genuchten et al., 2017). cles in the electrolyte, which could contribute to fouling and pas-
Oxyanions exhibit strong affinities to Fe(0) and Fe-oxyhydr(oxides), sivation (Alkhatib et al., 2020).
whereas Cl− and Br− ions replace oxyanions in the inner molecular Sulfide species can exist in certain types of effluents, such as
layer close to the electrode surface, effectively weakening the sur- produced water from oil production (Chow and Pham, 2019). The
face layer—which ultimately leads to pitting corrosion because Cl− presence of sulfides in Fe-EC is critical to acknowledge due to its
and Br− form highly soluble complexes with iron. Maintaining suf- reactivity within the electrochemical stability range of water, and
ficient availability of aggressive ions can be used to prevent elec- its ability to affect EC performance (Vepsäläinen, 2012). Sulfide will
trode passivation, a strategy that is discussed in Section 3.1. be present in different species depending on pH and DO concen-
The effect of passivation in Fe-EC is particularly evident in tration. H2 S is a weak acid (pKa = 7) and exists in acidic environ-
NO3 − solutions (Fig. 7). The passivating behavior of NO3 − is differ- ments, whereas HS− is present in alkaline conditions (Guan et al.,
ent than other ions because it does not form a strong interaction 2017). While the impact of sulfides on EC performance has not

7
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Table 2
Compilation of reported faradaic efficiencies, experimental conditions, and conclusions.

Electrode Faradaic
Material Sample Type Current Efficiency (FE) Conclusions Reference

Fe Cr3+ (0.03 M), NaCl 10.84 mA cm−2 64.5, 91.7% Higher FE in monopolar (Golder et al., 2007)
(0.05 M) in DI H2 O (Monopolar); configuration
32.52 (Bipolar)
mA cm−2
Fe Cr(VI), NO3 − in DI H2 O 0.16, 3.23 mA 1.7–111% ~100 FE at low current, no/low (Heidmann and
cm−2 FE at high current Calmano, 2008)
Fe Synthetic tap water; 0.86, 1.72, 3.45, 57–240% <pH 6: high FE (Sasson et al., 2009)
NaCl, organic buffers 6.9 mA cm−2 >pH 8: low FE
(MOPS, Glycine)
Fe Synthetic groundwater 2.8–44.4 mA 85–111% Multiple operating cycles: (Lakshmanan et al.,
(As) cm−2 high FE in first cycles; 2009)
progressively lower FE with
increasing cycles.
Fe Paper mill waste with 3.6–17.9 mA ~100% High FE in presence of sulfide (Vepsalainen et al.,
sulfide cm−2 2011)
Fe K2 Cr2 O7 , NaCl, NaNO3 , 4.36–30.74 mA 87.3–105.5%, Higher FE in presence of NaCl. (Golder et al., 2011)
Na2 SO4 , KH2 PO4 cm−2 110–210% (pH Higher FE at lower current
2–3) densities. Chemical dissolution
of Fe(0) in pH 2–3
Fe Na2 SO4 , NaCl in H2 O 5–125 mA cm−2 15–93% Lower FE at higher current (Dubrawski et al.,
density. Increased FE with 2015)
NaCl
Fe Synthetic groundwater 0.25–50 mA 0–100% FE ~0 in SO4 2− , CO3 2− , PO4 3− , (van Genuchten et al.,
cm−2 NO3 − , citrate, formate. ~100% 2017)
in soln with Cl− , Br−
Fe Synthetic smelting 4–10 mA cm−2 ~70–95% Lower FE at higher current (Xu et al., 2017)
wastewater with densities
cadmium, zinc,
manganese
Fe Synthetic mining 30 mA cm−2 77% Low FE caused by O2 evolution (Mamelkina et al.,
water (Cl− , NO3 − , 2017)
NH4 + , Cu2+ , Ni2+ ,
Zn2+ )
Fe Synthetic groundwater 0.8–10 mA cm−2 70–95% Efficiency high when CLR > (Müller et al., 2019)
4 min, and without oxyanions
Fe Groundwater 150A 12–35% Cell Potential: 7.8–11.8 V (Bandaru et al., 2020)
(before cleaning electrodes),
6.2–9.9 V (after cleaning)
Al Tap Water 2, 3, 5 mA cm−2 115–138% — (Jiang et al., 2002)
Al F− , Cl− , NO3 − , SO4 2− is 8.6 & 5.56 mA 24.3–128.2% Low FE with SO4 2− ; High FE (Hu et al., 2003)
Water cm−2 with Cl− & NO3 −
Al NaCl, Na2 SO4 in Water 0.5 mA cm−2 200–300% High Al3+ dissolution in >pH (Cañizares et al., 2005)
11
Al Na2 SO4 , NaCl, Ni2+ , 5–30 mA cm−2 120–224% High FE at lower i. Significant (Mouedhen et al.,
Cu2+ , Zn2+ chemical Al(0) dissolution at 2008)
cathode
Al Synthetic leachate 6V 0–131.1% No Al3+ dissolution in SO4 2− (Huang et al., 2009)
with NO3 − , SO4 2− , Cl− solutions
Al, Fe Commercial cutting oil 10 mA cm−2 Al: 150–300% Low FE in Fe-EC with NO3 − (Izquierdo et al., 2010)
in H2 O w/ NaCl, KCl, Fe: 0–100% salts & K2 SO4 2− (not Nas SO4 )
NH4 Cl, Na2 SO4 , K2 SO4
NaNO3 , KNO3
Al Algae in NaCl 2 mA cm−2 90–200% Increased FE with higher NaCl (Gao et al., 2010)
(0–0.47 g L−1 ) Water conc.
−2
Al NaCl, Na2 SO4 , 10 mA cm 100–133% — (Mansouri et al., 2011)
NaH2 PO4 in DI Water
Al (Pt CE) NaCl, 5 mA cm−2 100–180% FE generally ~100%; FE ~150% (Mechelhoff et al.,
Na2 SO4 , humic acid in in the presence of humic acid 2013b)
H2 O (>20 mg L−1 )
Al, Fe Acidogenic digestate, 9.3–27.8 mA 97–157% FE ~100% for iron, >100% for (Fayad et al., 2017)
(high organics conc.) cm−2 aluminum
pH 6.3
Al NaCl in DI H2 O (w/ 5 mA cm−2 145–320% Higher FE in NaCl (1 g L−1 ) (Ben Grich et al., 2019)
F− ), tap water (w/ F− ) than in tap water; Higher
Al(0) chemical corrosion at
cathode w/ increasing
temperature
Al Cl− , Fe2+ , Zn2+ , HCO3 − , 1.5 mA cm−2 130–270% Low FE in the presence of (Doggaz et al., 2019)
Ca2+ HCO3 −
Al 2,000 mg L−1 NaCl, pH 4–52 mA cm−2 80–116% Low FE at higher current (Chen et al., 2020)
8 densities
Ti 0.5 M NaCl in DI H2 O 7.5 mA cm−2 Slightly less Only electrochemical corrosion (El-Ghenymy et al.,
than 100% of Ti (no chemical) 2020)

8
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

been extensively studied, it has been shown in potentiodynamic


polarization studies that their presence effectively mitigates passi-
vation and favors pitting corrosion (Hernández-Espejel et al., 2011).
The same study used thermodynamic modeling to show that Fe(0)
reacts with sulfide to form mackinawite (FeS), pyrite (FeS2 ), and
subsequently Fe2 O3 /Fe3 O4 in later oxidation stages. A sequence of
chemical and electrochemical reactions was proposed as mecha-
nisms for these transformations: Initially, Fe(II) is electrochemically
produced at the anode, which forms FeS upon reacting with H2 S in
the acidic boundary layer (Eq. (13)). Chemically generated FeS ei-
ther precipitates on the electrode surface or is transported to the
bulk solution. The formation of FeS by this mechanism has been
reported elsewhere (Guan et al., 2017; Rickard and Luther, 2007;
Vepsäläinen, 2012). Furthermore, XRD analysis has shown that a
mixture of iron sulfides and Fe-oxyhydr(oxides) form on steel elec-
trodes after anodic polarization. It was therefore proposed that
FeS is further oxidized to Fe-oxyhydr(oxides) at the anode surface
(Eqs. (14) & (15)) (Hernández-Espejel et al., 2011). It was also hy-
pothesized that sulfur-species could oxidize to SO4 2− at high over-
potentials.
Fig. 8. Anodic (electrochemical) and cathodic (chemical) dissolution of aluminum
electrodes (anode and cathode), 1 g L−1 NaCl, i = 40 mA cm−2 . Reprinted by per-
2+ +
Fe + H2 S → FeS + 2H (13) mission from Springer Nature Customer Service Centre GmbH: Springer Journal of
Solid State Electrochemistry, Electrochemical dissolution of aluminium in electroco-
agulation experiments, Fekete et al., Copyright 2016.
2FeS + 3H2 O → Fe2 O3 + 2H2 S + 2H+ + 2e− (14)

3FeS + 4H2 O → Fe3 O4 + 3H2 S + 2H+ + 2e− (15) chemical dissolution of Al(0) are shown in Eqs. (16)–(18), respec-
tively.
Post-experiment transmission electron microscopy (TEM) im-
ages of electrodes have also revealed signs of elemental sulfur Al(s ) ↔ Al3+ (aq ) + 3e− E o = −1.66 V vs SHE (16)
formation as a result of H2 S oxidation (Chow and Pham, 2019).
Lastly, in an attempt to electrodeposit FeS2 on steel electrodes,
6H+ + 2Al(s ) → 3H2 + 2Al3+ (aq ) (17)
researchers found that the surface layer was primarily comprised
of crystalline FeS, whereas no FeS2 was observed. This was ex-

plained by FeS2 formation exhibiting sluggish reaction kinetics, 6H2 O + 2OH− + 2Al(s ) → 2[Al(OH )4 ] + 3H2 (18)
even though it was thermodynamically favorable under the exper-
imental conditions (Guan et al., 2017). When aluminum metal is exposed to air, a film of alumina
(Al2 O3 ) forms on the surface, which is stable in water at circum-
neutral pH (Badawy et al., 1999; Schmuki, 2002; Trompette and
2.3.2. Passivation behavior of aluminum electrodes in EC Vergnes, 2009). The oxide film thickness has been reported to
Aluminum ions only have one oxidation state (z = 3), but be between 1 and 5 nm using X-ray photoelectron spectroscopy
upon electrochemical dissolution, these ions can hydrolyze into (XPS) and electrochemical impedance spectroscopy (EIS) analyses
many types of polymeric hydroxide species (Mouedhen et al., (Martin et al., 2005; Mechelhoff et al., 2013a). Contrary to the
2008). Following Al(0) dissolution, Al(OH)2+ and Al(OH)2 + tend formation of oxyhydr(oxides) on iron electrodes, the Al2 O3 film
to form in acidic conditions, whereas Al(OH)4− forms in al- is insulating, hence functioning as a dielectric (Schmuki, 2002).
kaline media (Phalakornkule et al., 2010). Meanwhile, solid The insulating Al2 O3 surface layer must be broken down to facili-
Al(OH)3 /Al2 O3 exist within the range of pH 4–9.5 (Moussa et al., tate effective EC performance and is replaced by a porous layer of
2017; Roberge, 2008). Interestingly, observed concentrations Al(OH)3 /Al2 O3 during EC operation (Badawy et al., 1999; Gao et al.,
of aluminum in the electrolyte often exceed that of 100% 2010; Mansouri et al., 2011). Furthermore, on the aluminum Pour-
faradaic efficiency (Cañizares et al., 2007, 2005; Chen et al., baix diagram (Fig. 9), it is seen that Al(0)→Al(OH)3 oxidation oc-
20 0 0; Izquierdo et al., 2010; Mouedhen et al., 2008). It has curs in circumneutral pH, which is another example of a solid
been shown that Al(III) is primarily produced electrochemi- phase transformation that contributes to electrode surface layer
cally at the anode, but the “super-faradaic” behavior is a re- growth.
sult of chemical corrosion at both the anode and the cath- Passivation has been studied for aluminum electrodes with dif-
ode (Fig. 8) (Fekete et al., 2016), where the cathodic chemi- ferent electrolyte compositions. Solutions containing Cl− , F− , and
cal corrosion of Al(0) is catalyzed by the alkaline environment NO3 − have shown to decrease electrochemical overpotentials and
(Mouedhen et al., 2008). produce higher faradaic efficiencies, compared to solutions with
Without applying an electrical current, Al(0) dissolution in- high concentrations of SO4 2 − (Hu et al., 2003; Mansouri et al.,
creases dramatically above pH 11, compared to iron, which cor- 2011). The passivating effect of SO4 2 − in Al-EC has also been
rodes spontaneously below pH 4 (Cañizares et al., 2007). It has reported elsewhere (Huang et al., 2009; Izquierdo et al., 2010;
also been reported that the chemical corrosion of Al(0) at the Mansouri et al., 2011), and it is well established in the cor-
cathode accelerates with increasing temperature (Ben Grich et al., rosion literature (Lee and Pyun, 20 0 0; Pyun et al., 1999). Po-
2019). Meanwhile, the authors claimed that the electrochemical tentiodynamic polarization studies have shown that pitting po-
dissolution of the anode was unaffected by temperature changes. tentials are increased, and current densities are decreased in
Furthermore, it has also been shown (using gas chromatography) SO4 2 − and H2 PO4 − solutions compared to in high concentration
that Al(0) is oxidized by H+ , hence generating H2(g) at the anode NaCl electrolytes (Mansouri et al., 2011). In the presence of Cl− ,
(Fuladpanjeh-Hojaghan et al., 2019). The electrochemical and the the Al(OH)3 /Al2 O3 passivation layer breaks down according to

9
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

tion of Al(0) (Eq. (18)). However, a thick Al(OH)3 surface layer was
observed on the anode in the presence of NO3 − , compared to no
observable oxide film in the Cl− electrolyte (Hu et al., 2003). Fur-
thermore, the reduction mechanisms for NO3 − is corroborated by
relatively higher pH values observed in NO3 − solutions than in Cl−
electrolytes (Huang et al., 2009). The authors hypothesized that the
lower pH in the presence of Cl− was caused by the anodic oxi-
dation of Cl− →Cl2(g) and further production of HOCl that had a
buffering effect, which enhanced the speciation of Al-hydroxides.
Consequently, higher contaminant removal efficiencies were re-
ported in the Cl− solutions compared to in the NO3 − electrolytes,
even though the faradaic efficiencies were similar in both cases.
The buffering effect of high Cl− concentrations has been reported
elsewhere (Ben Grich et al., 2019).

6NO−
3 + 18H2 O + 10Al → 10Al (aq ) + 3N2 + 36OH
0 3+ −
(24)

Fig. 9. Pourbaix diagram of aluminum in water at 25 °C. The dashed lines represent
the electrochemical stability range of water. Produced based on calculations in the 3NO− 0
3 + 18H2 O + 8Al → 8Al
3+
+ 3NH3 + 27OH− (25)
MEDUSA-HYDRA software (Puigdomenech, 2019).

3NO−
3 + 3H2 O + 2Al (s ) → 2Al
3+
+ 3NO−
2 + 6OH

(26)

Eqs. (19)–(23) (Mansouri et al., 2011), which is conducive to pit- It has also been reported that the degree of passivation is de-
ting corrosion. pendent on electrode roughness, i.e., depassivation is not as likely
to occur on smooth electrodes due to enhanced structural in-
Al2 O3 + 6Cl− + 6H+ → 2AlCl3 + 3H2 O (19) tegrity of the surface layers (Mechelhoff et al., 2013a). However,
it is not apparent if this phenomenon is unique to the nature of
Al(OH)3 /Al2 O3 surface layers, or if it applies to Fe-EC passivation
Al(OH )3 + Cl− → Al(OH )2 Cl + OH− (20)
as well.
As mentioned in Section 2.1. formation of AlPO4 on electrode
Al(OH )2 Cl + Cl− → Al(OH )Cl2 + OH− (21) surfaces has been observed in phosphate electrolytes, which rein-
forces the passivating surface layer (Mansouri et al., 2011). Also,
Al(OH )Cl2 + Cl− → AlCl3 + OH− (22) the presence of carbonates, phosphates, silicates, and arsenates has
shown to lower F− removal in Al-EC while operating galvanostati-

AlCl3 + Cl− → AlCl4 (23) cally (Vasudevan et al., 2011). However, it was unclear if this was
a result of electrode passivation. Instead, it was suggested that car-
At sufficiently high SO4 2 − concentrations, it was observed that bonates passivated the anode surface, whereas the other ions de-
no Al(0) dissolved at all while operating at a constant voltage creased F− removal by competitive adsorption on the aluminum
(Huang et al., 2009). Additionally, the electrolyte pH remained the coagulant. However, as mentioned in Section 2.2., F− has been
same throughout the experiment, indicating that little or no cur- shown to passivate aluminum electrodes in EC (Hu et al., 2003).
rent passed through the reactor due to extreme passivation. Other Furthermore, the passivating effects of carbonates and phosphates
studies have reported evidence of current termination during po- have also been reported elsewhere (Doggaz et al., 2019; Dura and
tentiostatic operation in SO4 2 − electrolytes, e.g., an aluminum an- Breslin, 2019). Lastly, the effect of organics such as humic acids
ode surface showed little sign of corrosion in post-experiment on the anodic dissolution and passivation of Al(0) is not appar-
scanning electron microscopy (SEM) images (Hu et al., 2003). There ent (Mechelhoff et al., 2013a, 2013b), but it has been reported
was no sign of a thick or porous layer, indicating that the surface that higher concentrations of humic acids form a gelatinous layer
layer was thin and not visible in SEM. In contrast, other researchers on the anode surface, which effectively inhibits metal dissolution
showed that Al(0) dissolution was unaffected by changing the sup- (Yildiz et al., 2008).
porting electrolyte from NaCl to Na2 SO4 (Cañizares et al., 2005). In
this case, the authors concluded that metal dissolution was more 2.3.3. Summary of passivation of different electrodes
sensitive to pH changes, i.e., higher alkalinity favoring Al(0)→Al(III) The presence of most oxyanions in the electrolyte is detrimental
dissolution. However, the salt concentrations used in these experi- to both Al- and Fe-EC performance. The exception is that NO3 − is
ments were 2.45 and 3.0 g L−1 , which are high. Side reactions such passivating in Fe-EC (Izquierdo et al., 2010; Thiam et al., 2014), but
as Cl2(g) evolution, triggered by high Cl− concentrations, could ex- possibly conducive to chemical corrosion in Al-EC (Hu et al., 2003).
plain these trends. In contrast, some authors have recommended Few studies have investigated the effect of cations on EC perfor-
using Na2 SO4 over NaCl as a supporting electrolyte in Al-EC due mance, but Trompette and Vergnes, 2009 found that the pres-
to the enhanced conductivity of SO4 2− (Yildiz et al., 2008). The ence of Na+ had a negligible impact on Al-EC operation, whereas
choice between different supporting electrolytes in EC has been Izquierdo et al., 2010 observed that K+ had a detrimental effect
a topic of controversy, and Ghernaout and Ghernaout, 2011 pub- on faradaic efficiencies in Fe-EC. Furthermore, bench-scale experi-
lished a review paper on the contradicting conclusions that have ments demonstrated that electrode fouling was not as pronounced
been reported. in Al-EC as in Fe-EC (Shamaei et al., 2018). However, severe mate-
While NO3 − is detrimental to Fe-EC performance, it may pro- rial precipitation on the electrodes in Al-EC has been reported on
mote metal dissolution in Al-EC (Hu et al., 2003). This effect may the pilot-scale (Schulz et al., 2009). It is also unclear if there are
be caused by NO3 − reduction to NO2 − , NH3 , and N2 at the cathode differences in passivation behavior for electrodes made of differ-
(Eqs. (24)-(26), thus oxidizing Al(0)→Al(III) and generating OH− ent grades of steel and aluminum, which should be investigated in
(Hu et al., 2003). The increased concentration of OH− could trig- future studies. Interestingly, Müller et al., 2019 found that overpo-
ger a feedback loop that further promotes the chemical dissolu- tentials were higher when using lower purity steel in Fe-EC field

10
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

trials, compared to when using high purity steel in laboratory ex- sampling the bulk solution followed by acid digestion and ICP-OES
periments. However, it is unclear if the electrode purity was the analysis (Müller et al., 2019). The potential issue is that the total
main contributor to this effect. Lastly, while SO4 2 − ions are pas- electrochemically dissolved iron is distributed between the bulk
sivating for both materials, sulfate electrolytes have shown to fa- solution, the electrode surface layer, and the precipitated sludge.
cilitate anodic dissolution in Ti-EC (El-Ghenymy et al., 2020). This Ultimately, efficient coagulant mixing is essential, and high Fe-
observation highlights the complex nature of electrode passiva- oxyhydr(oxide) concentrations in the electrolyte are imperative to
tion/dissolution, which depends upon both the electrode material sustaining high treatment performance.
and the electrolyte composition, which can vary widely. Recent studies have provided more nuanced insights into the
cause of low faradaic efficiencies in Fe-EC (Bandaru et al., 2020;
2.4. Passivation and faradaic efficiency Müller et al., 2019). For example, it has been demonstrated that
once a surface layer has formed, electrochemically produced Fe(II)
It is generally presumed that higher faradaic efficiencies are is trapped within the lattice of the surface layer (Bandaru et al.,
correlated with enhanced contaminant removal, and maintaining 2020; Müller et al., 2019; van Genuchten et al., 2017, 2016). Con-
high efficiency is imperative to sustaining efficient EC performance sequently, dissolved metal is hindered from diffusing or migrating
(Müller et al., 2019). The faradaic efficiency of a given EC pro- to the bulk solution, hence accelerating the surface layer buildup
cess is related to the degree of electrode passivation, and as men- (Fig. 10). Due to the porous and conductive nature of this type of
tioned previously, severe surface layer growth can adversely af- surface layer, it is non-passive, which has been shown in chronopo-
fect faradaic efficiency by triggering side reactions or terminating tentiometric studies (Müller et al., 2019). Therefore, assuming that
the current. This is particularly evident in experiments with high this effect is responsible for declines in electrochemically dissolved
concentrations of oxyanions in the electrolyte (Hu et al., 2003; metal, faradaic efficiencies of 100% would be expected if they were
Huang et al., 2009; Izquierdo et al., 2010; Müller et al., 2019). determined from lost electrode mass after removing the surface
Therefore, the relationship between the faradaic efficiency and pas- layer post-experiment. This method has been used by one study
sivation is important for understanding the electrochemical effi- (Cesar Lopes Geraldino et al., 2016), which still observed variations
cacy of the EC process. in the faradaic efficiency of electrode consumption under different
Generally, faradaic efficiencies are around 100% in Fe-EC, experimental conditions. However, the connection between coagu-
independent of current density and the DO concentration lant accumulation in the surface layer, faradaic efficiency, and op-
(Lakshmanan et al., 2009). Low faradaic efficiencies have been ob- erating conditions merits further investigation.
served at high current densities, which has been attributed to Alternatively, a shift from Fe(0)→Fe(II) to Fe(0)→Fe(III)
unwanted anodic O2(g) evolution that was corroborated by bub- (Eq. (27)) at higher anodic overpotentials is thermodynami-
bles forming on the anode (Dubrawski et al., 2015). Additionally, cally favorable over O2(g) evolution and could make it seem like
the faradaic efficiency dropped from 116% to around 80% in Al-EC the observed faradaic efficiency is low (Moreno et al., 2009). As-
when increasing the current density from 18 mA cm−2 to 52 mA suming that this phenomenon takes place, there would be a shift
cm−2 (Chen et al., 2020). Low faradaic efficiencies have also been in the molar rate of electrochemically dissolved metal because the
reported for experiments carried out at pH 9, possibly caused charge transfer number would increase from 2 to 3. Therefore, less
by O2(g) evolution, which is more favorable in alkaline conditions iron coagulant would be produced per unit charge, which would
(Sasson et al., 2009). As mentioned in Section 2.3.1, an increased lower the calculated faradaic efficiency, assuming a two-electron
current density (Eq. (12)) strengthens the anodic polarization, transfer reaction (Eq. (5)). However, it has been suggested that
which attracts particles that could contribute to electrode passi- direct Fe(0)→Fe(III) oxidation is negligible (Sasson et al., 2009),
vation. Additionally, once O2(g) evolution is initiated, γ -FeOOH for- and none of the reviewed studies reported evidence of this oc-
mation is favored, which further promotes passivation. The forma- curring. In addition, using electrodes of different steel grades
tion of O2(g) has either been observed or assumed to be the cause could also affect faradaic efficiency. For example, impurities (Ni, V,
of faradaic efficiencies lower than 100% (Cañizares et al., 2005; Cr) could be preferentially oxidized over Fe(0). Few studies have
Gao et al., 2010; Golder et al., 2007; Heidmann and Calmano, 2008; compared using electrodes of different steel/aluminum purities,
Kobya et al., 2006; Mamelkina et al., 2017; Mollah et al., 2004b; but Müller et al., 2019 found that using a higher purity iron elec-
Müller et al., 2019; van Genuchten et al., 2017; Vepsäläinen, 2012). trode did not enhance the observed faradaic efficiency in Fe-EC.
The most convincing evidence substantiating O2(g) generation are Nevertheless, using low purity metal may be problematic due to
DO concentration measurements exceeding saturation values in Fe- the release of toxic impurities into the water.
EC in a NO3 − electrolyte operating at 10 mA cm−2 (van Genuchten
Fe(s ) ↔ Fe3+ (aq ) + 3e− E o = −0.037 V vs SHE (27)
et al., 2017). As mentioned in Section 2.3.2., it was also reported
and confirmed using confocal imaging and gas chromatography Another suggested hypothesis is that Fe(II) and Fe(III) ions are
that H2(g) forms on the anode—not O2(g) —in Al-EC when oper- reduced to Fe(0) at the cathode, which would decrease the mea-
ating with current densities up to 16 mA cm−2 (Fuladpanjeh- sured faradaic efficiency (van Genuchten et al., 2016; Yang et al.,
Hojaghan et al., 2019). Others have also hypothesized or observed 2015). However, this hypothesis has not been thoroughly inves-
H2(g) formation in Fe-EC due to chemical corrosion (Eq. (11)) tigated. Moreover, the evolution of Br2(g) and Cl2(g) is thermody-
(Cañizares et al., 2007; Golder et al., 2011; Guan et al., 2017). namically possible at sufficiently high electrode potentials, and oc-
Furthermore, it has been suggested that O2(g) evolution does not curs in electro-oxidation with boron doped diamond electrodes
take place in Al-EC due to the thermodynamic favorability of (Ferro et al., 20 0 0). However, it has been suggested that these re-
Al(0)→Al(III) dissolution (Mouedhen et al., 2008). Therefore, the actions are unlikely to occur in EC, even at overpotentials above
extent to which anodic O2(g) evolution in EC is responsible for their standard reduction potentials (van Genuchten et al., 2017).
observed faradaic efficiencies below 100% is still uncertain. Other Direct-oxidation of organics at the anode surface could also cause a
mechanisms that can affect coagulant production must be ex- drop in the measured faradaic efficiency, while positively affecting
plored. treatment performance. However, direct electrochemical oxidation
During EC operation, measured metal dissolution quantities of organics requires large overpotentials, and there is no evidence
could be artificially low due to inhomogeneous concentrations that these processes occur in EC (Ghernaout, 2013).
within the reactor (Bandaru et al., 2020). As mentioned in the The causes of low faradaic efficiencies observed in Fe-EC remain
introduction, the faradaic efficiency is commonly determined by unclear. However, many of the reviewed papers indicate that EC

11
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Fig. 10. Illustration of the surface layer buildup in Fe-EC, and consequent enmeshment of electrochemically dissolved metal, which impedes mass transfer to the bulk
solution. Reprinted from Chemosphere 153, van Genuchten et al., Formation of macroscopic surface layers on Fe(0) electrocoagulation electrodes during an extended field
trial of arsenic treatment, 270–279, 2016, with permission from Elsevier.

performance is heavily affected by the buildup of a passivating sur- As described previously, the addition of an aggressive ion such as
face layer on the anode. This phenomenon consequently increases Cl− triggers pitting corrosion, which breaks down the surface oxide
the anodic overpotential and may change the nature of the electro- layer.
chemical activity at the anode. By inspecting the faradaic efficien- It has been shown empirically that the pitting potential is pro-
cies reported by studies (Table 2), it is evident that low efficiencies portional to the logarithmic ratio of the concentrations of the ag-
are more frequently observed in Fe-EC than in Al-EC. Therefore, gressive and inhibiting ions in the electrolyte (Lumsden et al.,
sustaining efficient coagulant dosing is seemingly more attainable 1981; Refaey, 2005). The linearized expression for this relationship
in Al-EC, which is likely caused by the thermodynamic favorabil- is shown in Eq. (28). Experiments at different electrolyte concen-
ity of Al(0)→Al(III) oxidation (Eq. (16)). Izquierdo et al., 2010 also trations have been fitted to this model (Dura and Breslin, 2019;
reached this conclusion, i.e., Al-EC was more effective than Fe-EC Mouedhen et al., 2008).
at avoiding passivation and maintaining effective electrode disso-
Cx
lution under identical operating conditions. Consequently, this in- E p = A + Blog (28)
dication gives Al-EC a distinctive advantage over Fe-EC in long I
term continuous EC operation. However, a systematic study of Where constants A and B are experimentally determined, and Cx
long term Al-EC, similar to those for Fe-EC (Amrose et al., 2014; and I are the concentrations of the aggressive and the inhibitor
Bandaru et al., 2020; Müller et al., 2019), has not been reported. As ions, respectively. In Fe-EC operation, it has been suggested that
mentioned in Section 2.3.3., severe material precipitation on alu- 1 mM of NaCl should be added for every 20 and 100 mM of NO3 −
minum electrodes has been reported to occur during pilot-scale EC and HCO3 − in the solution, respectively—to mitigate passivation
applications (Schulz et al., 2009). and ensure high faradaic efficiency (Fig. 11) (van Genuchten et al.,
2017). In the presence of SO4 2− , it has been reported that main-
3. Passivation mitigation strategies taining a [Cl− ]/[SO4 2− ] ratio between 0.1–0.8 is sufficient to avoid
passivation and ensure high faradaic efficiency (Huang et al., 2009;
Researchers have tested a variety of techniques to remove elec- Mechelhoff et al., 2013a; Mouedhen et al., 2008; Trompette and
trode surface layers, or to avoid their formations. Implementing Vergnes, 2009).
effective strategies will enable continuous coagulant dosing while However, there are several drawbacks to adding aggressive ions
keeping the cell potential low during long-term EC operation. The to the electrolyte. Firstly, the added salts may need to be re-
depassivation and performance enhancement strategies that have moved in subsequent treatment steps before the treated water
been reported are examined in the following sections. could be re-used or discharged into the environment. Secondly,
a high Cl− concentration increases the risk of generating haz-
3.1. Aggressive ion addition ardous byproducts, especially in the presence of organics, which
could form organochlorinated compounds (Donaldson et al., 2002;
One of the most frequently recommended depassivation tech- Gotsi et al., 2005; Heidmann and Calmano, 2008; Kabdaşli et al.,
niques in EC is to add an aggressive ion to the electrolyte, such as 2009; Khandegar and Saroha, 2013). Furthermore, it has been
Cl− or Br− (Arroyo et al., 2009; Chellam and Sari, 2016; Dura and shown that adding too much Cl− can disrupt coagulation, e.g., dis-
Breslin, 2019; van Genuchten et al., 2017; Xu et al., 2014, 2015). solving Al(OH)3(s) to AlCl4 − (aq) (Wang et al., 2009). It was reported

12
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Fig. 11. Experimental data showing the abrupt decline in faradaic efficiency, coinciding with a spike in the anodic potential at specific electrolyte ratios. i = 1 mA cm−1 , pH
7.5. van Genuchten et al., 2017. Factors affecting the faradaic efficiency of Fe(0) electrocoagulation. J. Environ. Chem. Eng. 5, 4958–4968.

that adding > 1 g L−1 NaCl reduced the contaminant removal Vasudevan et al., 2011; Vasudevan and Lakshmi, 2012). AC-EC has
rate by preferential sorption of Cl− on the Al-coagulant (Abdel- also been tested with iron electrodes (Mansoorian et al., 2014),
Aziz et al., 2020). Additionally, it has been shown that Cl− in- aluminum (Cerqueira et al., 2014; Karamati-Niaragh et al., 2019;
creases the aqueous stability of certain dissolved contaminants and Vasudevan et al., 2011), zinc (Vasudevan and Lakshmi, 2012) and
metal-oxyhydr(oxides), effectively inhibiting the coagulation pro- magnesium electrodes (Kamaraj et al., 2013).
cess (Abdel-Aziz et al., 2020; Golder et al., 2011, 2007). Another Higher contaminant removal per unit mass of metal dissolved
concern is the increased sludge production associated with adding has been observed in AC-EC, possibly due to enhanced coagu-
salt to the water (Maha Lakshmi and Sivashanmugam, 2013). lant mass transfer caused by the corrosion of both electrodes
Hence, aggressive ion addition should be optimized to avoid both (Karamati-Niaragh et al., 2019). Ultimately, most studies have
electrode passivation and the undesirable side effects associated found that AC-EC is comparatively more energy efficient than DC-
with excessive ion concentrations (Maha Lakshmi and Sivashan- EC, which is often hypothesized to be a consequence of dimin-
mugam, 2013). ished electrode passivation (Alkhatib et al., 2020; Bian et al., 2019;
Table 3 provides an overview of the studies that have explored Kamaraj et al., 2013; Vasudevan et al., 2011; Vasudevan and Lak-
using aggressive ion addition as a depassivating strategy and shows shmi, 2012). The mechanisms responsible for avoiding or remov-
the suggested concentrations/ionic ratios that authors have recom- ing passivation in AC-EC have not been discussed in the literature.
mended. However, a proposed explanation for polarity reversal is provided
in Section 3.3.2., which can also be used to explain the mecha-
nisms for AC-EC.
3.2. Alternating current
Many authors have provided SEM images showing that
electrodes post-experiment display smoother surface morpholo-
Using current waveforms other than DC is an easily applicable
gies after AC-EC operation compared to after DC-EC (Fig. 12)
method that has been proposed to remove or inhibit the growth
(Kamaraj et al., 2013; Mansoorian et al., 2014; Vasudevan et al.,
of electrode surface layers in EC. Some studies have experimented
2011; Vasudevan and Lakshmi, 2012). Electrodes in DC-EC ex-
with using a sinusoidal alternating current (AC), which causes
hibit more crevices, localized corrosion, and material precipitation
the electrodes to oscillate between anodic and cathodic polarity
on the electrodes. Therefore, pitting corrosion is likely less pro-
(Cerqueira et al., 2014; Ghanizadeh et al., 2016; Kamaraj et al.,
nounced in AC-EC, primarily due to the cyclic interruption and
2013; Karamati-Niaragh et al., 2019; Mansoorian et al., 2014;
restarting of anodic dissolution, which leads to the continuous for-
Vasudevan et al., 2011; Vasudevan and Lakshmi, 2012). It has
mation of new points of localized corrosion. Consequently, this
been shown that AC-EC is more cost-effective than DC-EC for
leads to more uniform consumption of the anode, which could
various applications (Cerqueira et al., 2014; Kamaraj et al.,
overcome the problem of non-uniform electrode dissolution asso-
2013; Karamati-Niaragh et al., 2019; Mansoorian et al., 2014;

13
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Table 3
Reported results from studies using aggressive ion addition in EC.

Electrode
Material Sample Type Current Suggested Aggressive Ion Conc. Author

Fe NaCl, NaNO3 , Na2 SO4 , K2 Cr2 O7 10 mA cm−2 0.5 g L−1 NaCl (Lakshmipathiraj et al., 2008)
Fe Cr(VI) in NaCl & H2 SO4 DI Water 45 mA cm−2 0.7 g L−1 NaCl (Arroyo et al., 2009)
Fe Dyes, NaCl in H2 O 0.27–5.46 mA cm−2 1.5 g L−1 NaCl (Secula et al., 2013)
Fe K2 Cr2 O7 , NaCl, K2 SO4 in DI H2 O 10 mA cm−2 0.36–1.47 g L−1 (Xu et al., 2015)
Fe Synthetic Groundwater 0.25–50 mA cm−2 1 mM NaCl for every 20 mM NO3 − , (van Genuchten et al., 2017)
100 mM HCO3 −
Fe/Al Cr2 O7 2− in DI H2 O 15.3 mA cm−2 0.57 g L−1 NaCl, 0.285 g L −1 KCl (Keshmirizadeh et al., 2011)
Fe, Al Diluted Tanning Waste 20 (optimal), 40 mA cm−2 1 g L−1 NaCl (90% COD removal) (Maha Lakshmi and
Sivashanmugam, 2013)
Al Synthetic leachate with NO3 − , SO4 2− , 45 mA cm−2 [Cl− ]/[SO4 2− ] = 0.4 (Huang et al., 2009)
Cl−
Al Na2 SO4 , NaCl Ni2+ , Cu2+ , Zn2+ 0.5–23 mA cm−2 [Cl− ]/[SO4 2− ] = 0.15 (Mouedhen et al., 2008)
Al Laundry waste 1–7 V 0.25 g L−1 NaCl (Wang et al., 2009)
Al Milk with NH4 Cl, NaCl, NH4 SO4 , 8 V (MSE) WE [Cl− ]/[SO4 2− ] > 0.1 (Trompette and Vergnes, 2009)
Na2 SO4
Al Algae in NaCl Water 2 mA cm−2 0.29 g L−1 NaCl (Gao et al., 2010)
Al Na2 SO4 , NaCl, humic acid in pH 7 RO 0–1.5 mA cm−2 [Cl− ]/[SO4 2− ] = 0.8 (Mechelhoff et al., 2013a)
H2 O
Al Synthetic NO3 − solution; pH 3 - 11, 2.94–10.3 mA cm−2 1 g L−1 (Abdel-Aziz et al., 2020)
NaCl 1–2.5 g L− 1

Xu et al., 2019). However, AC-EC has shown to offer benefits


such as reduced sludge formation, lower energy consumption, im-
proved mass transfer characteristics, and delayed electrode pas-
sivation (Alkhatib et al., 2020; Cerqueira et al., 2014; Karamati-
Niaragh et al., 2019; Mansoorian et al., 2014; Xu et al., 2019). In
contrast, one study reported that DC-EC was better than AC-EC
for the removal of F− with both iron and aluminum electrodes
(Ghanizadeh et al., 2016). Yet, in total, there have only been a
handful of authors that have examined the effects of AC opera-
tion, and more research is required to develop a comprehensive
understanding of the mechanisms, benefits, and drawbacks of this
technique. The operating conditions reported by authors of AC-EC
studies are included in Table 4 below.

3.3. Polarity reversal

3.3.1. Principles of polarity reversal electrocoagulation


In contrast to using a sinusoidal current to invert electrode po-
larity, operating with polarity reversal entails periodically switch-
ing the direction of a direct current, thus resulting in a square
wave (Fig. 13a). This technique is the most commonly used pas-
sivation control strategy in EC, and is often referred to as an al-
ternating pulsed current (Eyvaz et al., 2009; Keshmirizadeh et al.,
2011; Mao et al., 2008; Yang et al., 2015) or as polarity re-
versal electrocoagulation (PR-EC) (Bao et al., 2020; Liu et al.,
2018; van Genuchten et al., 2017; Vasudevan et al., 2011). Some
researchers have operated PR-EC with intermittently terminat-
ing/restarting (i.e., pulsing) the current in between polarity rever-
sals (Fig. 13b), which was proposed to enhance the mass trans-
fer of the coagulant to the bulk solution (Asaithambi et al., 2016;
Hasani et al., 2019; Ren et al., 2011; Xin et al., 2018; Zhou et al.,
2020). PR-EC is relatively easy to implement, and the only ad-
ditional infrastructure required is a DC time relay, or manually
Fig. 12. SEM images of the aluminum anode post-experiment after (a) AC-EC and switching the positive/negative terminals. The idea of reversing po-
(b) DC-EC for operating with a F− electrolyte (5–20 mg L−1 ). Reprinted from Water
larity as a method to consume both electrodes can be found in
Sci. Technol. 65, Vasudevan and Lakshmi, Effect of alternating and direct current
in an electrocoagulation process on the removal of cadmium from water, 353–360, patents from the 1970s (Miller, 1977, 1978). PR-EC later received
Copyright 2011, with permission from Elsevier. attention in the 1980s’ scientific literature as an electrode clean-
ing technique (Grøterud and Smoczynski, 1986; Jianqian, 1988). It
has since then been established as a strategy to depassivate, sus-
ciated with small electrode gaps and high solution conductivities tain high treatment performance, and to extend electrode lifetime
(McBeath et al., 2020). (Mollah et al., 2001).
It has consistently been reported that faradaic efficiencies PR-EC has been evaluated for several applications (Table 4).
are lower in AC-EC than in DC-EC (Alkhatib et al., 2020; Most studies have operated PR-EC with current reversals tak-

14
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Table 4
Compilation of studies that have investigated the performance of PR-EC and AC-EC.

Electrode Current Energy


Material Type PRT Sample Type Current Consumption Conclusions Author

Fe PR 5–30 min Cr(VI) & NO3 − in DI 0.16, 3.23 mA 0.003–16.83 kWh Higher performance at low (Heidmann and
H2 O cm−2 m−3 current density. No Calmano, 2008)
reported effect of PRT
Fe PR 0.5–4.2 min Sea water 1.8–1.9 mA — Highest Farad. Eff. at (Timmes et al.,
cm−2 4 min 12 s PRT 2010)
−3
Fe PR 30 min K2 Cr2 O7 , KCL, NaOH, 10–40 46 kWh m No other PRT tested (Dermentzis et al.,
HCl mA cm−2 2011)
Fe PR 5 min Dyes, NaCl in H2 O 0.27–5.46 mA DC-EC: ~0.26 kWh PR-EC: lower energy (Secula et al.,
cm−2 m−3 ; consumption, same 2013)
PR-EC: ~0.23 kWh removal performance as
m−3 DC-EC
Fe PR Anode/cathode Arsenic in 45 A: 4.3–5 C 0.62–2.31 Mechanical cleaning (Amrose et al.,
reversed each groundwater L−1 min−1 kWh m−3 ; enhanced performance; 2014)
exp. cycle 0.83–1.04 effects of PR-EC not
US$ m−3 discussed.
Fe AC Not Reported Pb & Zn in battery 2–10 AC-EC: 0.69 kWh AC-EC higher performance (Mansoorian et al.,
water mA cm−2 m−3; than DC-EC 2014)
DC-EC: 1.97 kWh
m−3
Fe PR 0–15 min Electroplating effluent, 0.3–1.5 mA 10–20% lower 10–15 min PRT optimal for (Shao-qin et al.,
Cr(VI), pH 5–10 cm−2 energy Cr(VI) removal 2014)
consumption in
PR-EC than in
DC-EC
Fe PR 10 s - 30 min Cr(VI) & K2 SO4 in 10 — PRT >9 min 50 s required (Yang et al., 2015)
water mA cm−2 for high Farad. Eff.
Fe PR Anode/cathode Groundwater 45 A — PR and mech. cleaning of (van Genuchten
switched each electrodes can reduce et al., 2016)
exp. operating costs
Fe PR Dairy plant effluent, 1.5 A — Higher electrode (Cesar Lopes
Anodes/Cathodes pH 4.5 consumption in PR-EC Geraldino et al.,
switched each than DC-EC 2016)
exp. (60 min)
Fe PR 15 min Distillery effluent, pH 1.3 1.3–1.32 Improved performance in (Asaithambi et al.,
3–11 mA cm−2 kWh m−3 PR-EC compared to in 2016)
DC-EC
Fe PR 52.5 min & 21 Synthetic Groundwater 0.8 PR-EC: Increased electrode fouling (Müller et al.,
min mA cm−2 0.21–0.22 kWh in PR-EC caused by 2019)
m−3 ; trapping of Fe(II) in CaCO3
DC-EC:
0.2–1.48 kWh m−3
Fe PR 30, 120, 240 s Estrogen, Na2 SO4 , 4.16, 8.3, 16.7 — 30 s optimal PRT for (Maher et al.,
NaHCO3 in DI H2 O mA cm−2 estrogen removal. 2019)
Fe AC 50 Hz K2CrO4, NaCl, Ethanol, 0.27 AC-EC: Less electrode (Xu et al., 2019)
Diphenylcarbohy- mA cm−2 0.454 kWh m−3 ; consumption and better
drazide in DI DC-EC: removal for AC-EC than
H2 O 1.002 kWh m−3 DC-EC
(~96–97% Cr(VI)
removal)
PR 5 min Raw coke water 20 mA cm−2 1.93–3.30 PR-EC: removal efficiency (Ozyonar and
Fe + H2 O2 kWh m−3 than DC-EC Karagozoglu, 2015)
Fe, Al PR 4 s – 17 min 20 mg L−1 P 0.05 A 27 kWh m− 3 256 s optimal PRT for P (Grøterud and
removal Smoczynski, 1986)
Fe, Al PR 4 min Cr(VI), + ions in water 2–23 4–58 kWh m−3 PR-EC: decreased
mA cm−2 passivation, less energy (Keshmirizadeh et al.,
consumption than DC-EC 2011)
Fe, Al AC 50 Hz NaF (3–8 mg L− 1 ) 15–40 V — DC-EC: more effective F− (Ghanizadeh et al.,
removal than AC-EC 2016)
Fe, Al PR 2.5 min Brewery waste 3–30 Fe: 4.68 kWh m−3 PR-EC overall more (Eyvaz, 2016)
mA cm−2 (DC) efficient than DC-EC; Less
4.30 kWh m− 3 coagulant required per kg
(PR-EC) COD removed
Al: 4.85 kWh m− 3
(DC)
4.27 kWh m− 3
(PR-EC)
Fe, Al PR 0.01 s - 1 min Milk in tap water, pH 0.01–40 mA — PRT >1 min necessary for (Fekete et al.,
40 s 5.5. cm−2 high Farad. Eff. 2016)
NaCl, Na2 SO4 in DI
H2 O
Fe, Al PR 1, 3, 5 min Synthetic brine, pH 2.48–2.81 mA 0.63–2.89 No difference between PRT (Ashraf et al.,
9.44 cm−2 kWh m−3 & salt removal. Slight 2019)
increase in energy
consumption with 5 min
PRT
(continued on next page)
15
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Table 4 (continued)

Electrode Current PRT Sample Type Current Energy Conclusions Author


Material Type Consumption

Al PR Not Reported Greywater (cooking oil, 0.5–1.5 mA 30% lower energy Enhanced COD removal, (Mao et al., 2008)
scour, Na2 SO4 ) cm−2 consumption in and more uniform
PR-EC than DC-EC electrode dissolution in
PR-EC
Al PR 3–15 min Dyes in water 3.5–15 1.3–3.43 US$ (kg PR-EC: Decreased (Eyvaz et al., 2009)
(Disperse Yellow 241 mA cm−2 TOC)−1 passivation, enhanced TOC
and Reactive Yellow removal, lower cost than
135) DC-EC
Al AC 50 Hz NaF in DI H2 O 2.5–15 AC-EC: 1.883 kWh AC-EC prevents passivation (Vasudevan et al.,
mA cm−2 m− 3 ; 2011)
DC-EC: 2.541 kWh
m− 3
Al PR 1–30 s Methyl orange (MO), 150 PR-EC: 44 kWh (kg 15 s optimal PRT; PR-EC (Pi et al., 2014)
NaCl in H2 O mA cm−2 MO)−1 ; overall more effective than
DC-EC: 55 kWh DC-EC
(kg MO)−1
Al AC, PR 60 Hz (AC), Synthetic produced 20–100 mA 0.28 kWh m−3 Higher electrode (Cerqueira et al.,
5 min (PR) water (O/W emulsion) cm−2 consumption in PR-EC 2014)
than in AC-EC
Al PR 15, 30 min Groundwater (As) 7V 3.03 kWh m−3 No sign of passivation in (Mohora et al.,
(30 min PRT) PR-EC (6 h operation) 2014)
Al PR 5 min Produced water 1.2 V PR-EC + biochar PR-EC lower energy (Lobo et al., 2016)
filter: 0.15 kWh consumption than DC-EC
(kg SS)−1 for similar pollutant
removal
Al PR 4 min, 16 s Raw Sewage 10–15 A 0.53 kWh m−3 No other PRT tested (Smoczyński et al.,
2017)
Al PR 10 s Synthetic produced 89 mA cm−2 — No other PRT tested (Gobbi et al., 2018)
water
−3
Al PR 10 s PFOA in water 9V 0.04 kWh m No other PRT tested (Liu et al., 2018)
Al PR 0–3 min NaCl, NaHCO3 in DI 37.9 V 2.56 kWh m−3 3 min optimal PRT for (Wellner et al.,
H2 O (constant) depassivation and 2018)
coagulant production
Al AC 50 Hz NaNO3 & Na2 SO4 in 7–8 mA cm−2 DC-EC: US$54 (kg Electrode consumption (Karamati-
water NO3 − )−1 ; higher in AC-EC than in Niaragh et al.,
AC-EC: DC-EC 2019)
US$29 (kg NO3 − )−1
Al PR 0.5, 1 & 2 min Groundwater 3–10 0.92–1.95 1 min optimal PRT for F− (Betancor-
mA cm−2 kWh m−3 removal Abreu et al.,
2019)
Al PR, AC 0.1–4 s Synthetic bilge water 3.6–10.7 mA AC-EC: 0.977 kWh Less long-term passivation (Bian et al., 2019)
cm−2 m−3 ; & energy consumption in
DC-EC: 0.936 kWh AC-EC than in DC-EC
m− 3
Al PR 1–10 min Humic acids in DI H2 O 5–15 V DC-EC: 1.08 kWh Less electrode (Hasani et al.,
m−3 consumption and lower 2019)
PR-EC: 0.37 kWh passivation in PR-EC than
m− 3 in DC-EC for same removal
(~97–98% HA of humic acids
removal)
Al AC 50 Hz Secondary treated 0.8–4.3 mA — Better COD & TP removal (Alkhatib et al.,
effluent (WWTP), pH cm−2 performance in AC-EC than 2020)
6.9 DC-EC; Lower electrode
consumption in AC-EC
Zn AC 50 Hz Cd(NO3 )2 in DI H2 O 1–5 AC-EC: 0.655 kWh AC-EC prevents passivation (Vasudevan and
mA cm−2 m−3 ; Lakshmi, 2012)
DC-EC: 1.236 kWh
m−3
Mg AC 50 Hz CuNO3 in DI H2 O 0.1–10 AC-EC: 0.634 kWh AC-EC better performance (Kamaraj et al.,
mA cm−2 m−3 ; than DC-EC 2013)
DC-EC: 0.996 kWh
m− 3
Fe, Ti PR 12 s As(III/IV) groundwater 5–50 mA PR-EC: 0.101 kWh 6 s current break between (Xin et al., 2018)
(IrO2 /Ta2 O5 m−3 ; PR decreased fouling and
coating) DC-EC: 0.167 kWh passivation
m− 3
Mg, Al PR 2–8 min Dye (Indigo Carmine) 4.44, 7.30 mA — Short PRT favors dye (Donneys-
& NaCl in DI H2 O cm−2 removal, DC-EC favors Cl− Victoria et al.,
removal 2020)
Zn, Al PR 10 s PFSAs in synthetic & 6–12 V — No other PRT tested (Bao et al., 2020)
natural groundwater

16
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Fig. 13. Direct current with (a) polarity reversal, (b) pulsing polarity reversal (with current breaks).

ing place during operation (Eyvaz, 2016; Eyvaz et al., 2009; taminant removal per unit coagulant produced. It was also ob-
Gobbi et al., 2018; Keshmirizadeh et al., 2011; Liu et al., 2018; served that the coagulant formed in PR-EC had a larger surface
Mao et al., 2008; Müller et al., 2019; Ozyonar and Karago- area than the coagulant formed in DC-EC when operating with iron
zoglu, 2015; Timmes et al., 2010; Yang et al., 2015). Others have electrodes (Zhou et al., 2020).
switched the electrodes in between operating cycles (Amrose et al., Similar to AC-EC, PR-EC could mitigate the problem of non-
2014; Bandaru et al., 2020; Cesar Lopes Geraldino et al., 2016; uniform anodic dissolution, which causes pre-mature structural
Maha Lakshmi and Sivashanmugam, 2013; van Genuchten et al., failure of the electrodes (McBeath et al., 2020). Another benefit of
2016). PR-EC is that two electrode materials could be used to produce two
types of coagulants (Keshmirizadeh et al., 2011). For example, this
3.3.2. Benefits of polarity reversal electrocoagulation permits using iron and aluminum electrodes together, which could
Some authors have observed increased faradaic efficiencies in be beneficial for the treatment of certain types of wastes. For in-
PR-EC compared to in DC-EC for operating with both iron and alu- stance, it has been shown that Fe-EC is more effective at removing
minum electrodes (Cesar Lopes Geraldino et al., 2016; Ozyonar and sulfides, while Al-EC is better at abating silica in produced water
Karagozoglu, 2015; Wellner et al., 2018; Yang et al., 2015). This from in-situ bitumen extraction (Chow and Pham, 2019).
effect was explained by the inevitable passivation in DC-EC, thus The mechanism for how passivation is avoided or removed
triggering side reactions. The authors suggested that fouling is when operating AC-EC or PR-EC is not clearly understood. A pos-
avoided in PR-EC, hence sustaining effective metal dissolution. Ad- sible explanation could be scouring by the formation of H2(g) bub-
ditionally, SEM imaging and XPS analysis have shown that a less bles on the cathode (Bian et al., 2019; Timmes et al., 2010), which
pronounced electrode surface layer forms in PR-EC with aluminum mechanically remove the surface layer from the anode following
electrodes (Hasani et al., 2019; Mao et al., 2008). It was also polarity reversal. Also, calcium and magnesium minerals formed
demonstrated that PR-EC and DC-EC exhibit similar removal effi- on the cathode can dissolve in the acidic boundary layer after the
ciencies for current densities below 12 mA cm−2 for abating COD cathode becomes the anode. By changing electrode polarity, elec-
in brewery wastewater using aluminum electrodes (Eyvaz, 2016). trostatically adsorbed particles on the surface are repulsed, which
However, the electrodes in DC-EC started passivating when operat- also contributes to depassivation. The surface layer could also be
ing at 12 mA cm−2 , whereas passivation was not detected in PR-EC weakened by the dissolution of the underlying metal, causing it to
until current densities of 18 mA cm−2 . fracture (Bian et al., 2019). Understanding these mechanisms re-
It has been reported that PR-EC uses 30–60% less energy than quires further investigation.
DC-EC for similar contaminant removal efficiencies with aluminum
electrodes (Hasani et al., 2019; Mao et al., 2008). Meanwhile, en- 3.3.3. Effect of polarity reversal frequency
ergy consumption data reported for PR-EC with iron electrodes The polarity reversal time (PRT), i.e., the time in between
tend to be lower than for DC-EC, but the difference is not as dra- current reversals, has been reported to affect EC performance
matic as for aluminum (see Table 4). Apart from the alleged re- (Eyvaz et al., 2009; Grøterud and Smoczynski, 1986; Timmes et al.,
duced passivation, the cell potential in PR-EC is lower than in 2010; Yang et al., 2015). For example, short PRTs have shown to
DC-EC because H2(g) evolution is thermodynamically favorable in reduce treatment performance (Yang et al., 2015). Meanwhile, the
acidic pH. Therefore, cathodic H2(g) evolution will proceed at a nature of passivation becomes analogous to DC-EC when operat-
lower equilibrium potential following polarity reversal due to the ing with longer PRTs. These observations indicate that this variable
acidic environment at the cathode. However, this effect will dimin- can be optimized, but PRTs recommended in the literature differ
ish once the pH at the cathode interface equilibrates. (Table 4), which suggests that the optimal PRT is dependent on op-
Due to the electrochemical dissolution of all electrodes in PR- erating conditions such as electrolyte composition, current density,
EC, metal ions are added more uniformly, which increases mix- and electrode material. For example, it was concluded that a 5 min
ing and induces a higher degree of contaminant-coagulant inter- PRT was optimal for the treatment of organic dyes Disperse Yellow
actions. Consequently, the coagulation process becomes more effi- 241 and Reactive Yellow 135 in Al-EC (Eyvaz et al., 2009). Mean-
cient, which permits decreasing the charge loading rate—hence re- while, a 4 min PRT was recommended for removing phosphates
ducing sludge production (Hasani et al., 2019; Keshmirizadeh et al., (Al/Fe-EC) (Keshmirizadeh et al., 2011; Smoczyński et al., 2017), a
2011). It has also been suggested that PR-EC mitigates the effect of 15 s PRT for removing methyl orange dye (Al-EC) (Pi et al., 2014),
concentration polarization in Al-EC, i.e., enhancing the mass trans- 1 min for F− (Al-EC) (Betancor-Abreu et al., 2019), 15 min for Al-
fer of Al3+ and Al(OH)4 − that otherwise tend to concentrate at the EC operation (Chen, 2004), and 2.5 min for general EC operation
anode and cathode (respectively) during DC-EC operation (Pi et al., (Eyvaz, 2016). Additionally, a 10 s PRT has been used with alu-
2011, 2008). Consequently, the speciation of Al-hydroxide polymers minum electrodes to remove oil, perfluorooctanoic acids (PFOAs)
was improved in PR-EC, which is conducive to more efficient con- and perfluorosulfonic acids (PFSAs), although the results were not

17
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

compared to DC-EC, or other reversal frequencies (Bao et al., 2020; nomenon (Müller et al., 2019). Another possible explanation for the
Gobbi et al., 2018; Liu et al., 2018). relatively thicker surface layers in PR-EC could be the occurrence of
It has been observed that faradaic efficiencies vary based on changing pH boundary layers at the electrode interfaces. For exam-
PRT duration, particularly in Fe-EC. For example, it was reported ple, Fe(II/III) present in the acidic anodic boundary layer can form
that the faradaic efficiency increased from 45% to 70% when in- Fe-oxyhydr(oxides) in the alkaline conditions that develop follow-
creasing the PRT from 10 to 590 s, which consequently enhanced ing polarity reversal (Fig. 5). These species sorb to the electrode
Cr(VI) removal in Fe-EC (Yang et al., 2015). Low faradaic efficien- surface and contribute to the surface layer buildup, which impedes
cies associated with short PRTs in Fe-EC have also been reported Fe(II/III) mass transport to the bulk solution. Furthermore, allowing
by others (Ashraf et al., 2019; Fekete et al., 2016; Grøterud and for the pH boundary layers to re-establish after polarity reversal
Smoczynski, 1986; Timmes et al., 2010). As mentioned previously, requires a particular duration, which depends on operating con-
the bulk solution pH typically increases gradually during DC-EC ditions such as current density and electrolyte buffering capacity.
operation. However, it was observed that the pH increase is not Also, the slow dissolution kinetics of calcium and magnesium min-
as pronounced in PR-EC with short PRTs, compared to longer PRTs erals in the anodic acidic boundary layer adds additional time to
or DC-EC (Grøterud and Smoczynski, 1986). This phenomenon has achieve effective depassivation. Consequently, applying longer PRTs
been seen for other systems. For example, it took considerable favor the completion of these mechanisms.
time for the pH to regain its equilibrium value after reversing It has also been acknowledged that the “polarization time”,
polarity with aluminum electrodes in a continuous flow system i.e., the time it takes for the electrochemical reactions to
(Wellner et al., 2018). Low faradaic efficiencies with short PRTs are restart following polarity reversal, is detrimental to EC perfor-
consistent with findings from studies of AC-EC, which are equiv- mance (Grøterud and Smoczynski, 1986). Non-faradaic charg-
alent to PR-EC with very short PRTs (i.e., high-frequency reversal) ing/discharging of the capacitive electrochemical double layer
(Cerqueira et al., 2014; Karamati-Niaragh et al., 2019). In contrast, (Hasani et al., 2019) and faradaic side reactions could occur dur-
the removal of estrogen was reported to increase with shorter PRTs ing this time. Therefore, the expected anodic and cathodic reac-
when using iron electrodes (Maher et al., 2019). The authors con- tions (Eqs. (1) & (2)) are presumably inactivated during this pe-
cluded that this was a consequence of avoiding electrode passiva- riod (Bian et al., 2019). This proposition is corroborated by rela-
tion. However, an alternative explanation is that an elevated rate of tively lower pH and faradaic efficiencies observed in PR-EC com-
direct anodic oxidation caused the improved removal rates, rather pared to in DC-EC (Fekete et al., 2016; Grøterud and Smoczyn-
than enhanced coagulation in the electrolyte. By using PR-EC, di- ski, 1986). By approximating the duration of charging the capac-
rect oxidation takes place at both electrodes, and the rate of mass itive double layer in EC—assuming 1 cm electrode separation and
transfer of pollutants could be improved. Enhanced mass transfer C = 0.2 F m−2 (Conway, 2018)—periods within the range 2–100 ms
and electrode utilization have been observed when applying po- are obtained (based on electrolyte conductivities 1–20 mS cm−2 ,
larity reversal in the electro-oxidation process (Thostenson et al., where higher conductivity corresponds to faster charging). Most
2018). PR-EC applications operate within these electrolyte characteristics
Longer PRTs have also been tested for the treatment of ground- (Izquierdo et al., 2010; Shamaei et al., 2018; Yildiz et al., 2008),
water by Fe-EC (Müller et al., 2019). In this study, the faradaic effi- and must therefore account for capacitive charging during po-
ciency was lower for a PRT of 21 min than observed for switching larity reversal cycling. This effect may explain why no net cur-
polarity between every operating batch (105 min), which in turn rent was observed when operating PR-EC at 60 Hz (17 ms PRT)
had a lower faradaic efficiency than DC-EC. It was also found that (Bian et al., 2019). Additionally, AC-EC is often operated at 50 Hz
the anodic interfacial potential was lower in PR-EC than in DC- (20 ms PRT). Therefore, this strategy may become ineffective when
EC experiments, indicating that passivation was depressed. In con- treating low conductivity waste. This phenomenon may explain
trast, Cesar Lopes Geraldino et al., 2016 demonstrated that elec- why Ghanizadeh et al., 2016 concluded that AC-EC was ineffective
trode consumption was higher over 24 cycles (relative to DC-EC) at removing F− (using iron and aluminum electrodes) from syn-
when reversing polarity in between batches (60 min) in Fe-EC for thetic solutions with low electrolyte concentrations.
treating dairy plant effluent. The authors concluded that passiva- It is also possible to prompt electrochemical side reactions fol-
tion led to reduced coagulant production in DC-EC, which is plau- lowing polarity reversal, especially in Fe-EC. Fe-oxyhydr(oxides) are
sible due to the high organic content of the wastewater. The dif- either comprised of Fe(II), Fe(III), or a combination of both (as de-
ferences between the findings of these studies highlight the com- scribed in Section 2.3.1.). These species can be oxidized or reduced,
plex nature of PR-EC and the effect of electrolyte composition on given the right environmental conditions. For example, it is possi-
faradaic efficiency and treatment performance. Due to the problem ble that HFeO2 − (formed during a cathodic cycle in high pH) is
of passivation in EC, and the widespread belief that PR-EC may ad- oxidized to FeO2 − during the subsequent anodic period according
dress this issue, it is essential to consider the connection between to Eq. (29) Bard et al., 1985), which is thermodynamically favorable
performance, passivation, and PRT. over Fe(0)→Fe(II) oxidation (Eq. (9)). Fe(OH)2 formed at the cath-
ode can also oxidize to Fe3 O4 in the alkaline environment imme-
3.3.4. Mechanisms of passivation mitigation in polarity reversal diately following polarity reversal (Fig. 5). Additionally, Fe(III) and
EC—discussion Fe(II/III)-oxyhydr(oxides) formed in the anode vicinity can reduce
Few studies have attempted to provide mechanistic explana- to Fe(II) and FeO in a cathodic environment (Eqs. (30)-(34), which
tions for the association between PRT and PR-EC performance. As is thermodynamically favorable over H2(g) evolution (E° = 0 V vs
discussed in Section 2.4., the buildup of the electrode surface layer SHE [at pH 0]) (Gorski et al., 2016; Xu et al., 2019).
likely contributes to a decrease in faradaic efficiency by hinder-
ing coagulant mass transfer to the bulk solution (Bandaru et al., HFeO− − − − o
2 + OH ↔ FeO2 + H2 O + e E = −0.685 V vs SHE (29)
2020; Müller et al., 2019; van Genuchten et al., 2016). It has
been suggested that this effect is accelerated in PR-EC, which
Fe3+ + e− ↔ Fe2+ E o = 0.771 V vs SHE (30)
is supported by thicker electrode surface layers observed on
iron electrodes in PR-EC relative to DC-EC (Ashraf et al., 2019;
Müller et al., 2019). An enhanced trapping capacity of iron species Fe(OH )3 + e− ↔ Fe2+ + 3OH− (31)
by a CaCO3 surface layer that forms on the cathode before it
becomes the anode was proposed as the causation of this phe- Fe3 O4 + 2H+ + 2e− ↔ 3FeO + H2 O (32)

18
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

α -FeOOH + e− + 3H+ ↔ Fe2+ + 2H2 O E o = 0.768 V vs SHE (33)

α -Fe2 O3 + 2e− + 6H+ ↔ 2Fe2+ + 3H2 O E o = 0.769 V vs SHE (34)

As discussed in Section 2.4., it is also possible to reduce


Fe(II)→Fe(0) at moderate cathodic overpotentials (Fig. 5). For ex-
ample, this phenomenon has been utilized to make batteries
that electroplate Fe(II)→Fe(0) during charging at the cathode, fol-
lowed by oxidation back to Fe(II) during discharge (Hruska and
Savinell, 1981). This effect could also take place in EC, where an-
odically dissolved Fe(II/III) species could be reduced back to Fe(0)
following polarity reversal (Xu et al., 2019). Electroplated Fe(0)
could then be re-oxidized during the subsequent reversal cycle.
Therefore, over the entirety of the operating period, less metal is
transferred from the electrode to the bulk solution. For example,
it has been demonstrated using cyclic voltammetry that Fe(II/III)-
oxyhydr(oxides) are reduced/oxidized at the iron electrode follow-
ing anodic or cathodic current shifts (Yang et al., 2015). However, Fig. 14. Electrochemically dissolved metal with iron and aluminum electrodes at
varying PR frequencies, i.e., PRTs. i = 16 mA cm−2 , NaCl & NaSO4 electrolyte.
the authors did not consider how this would affect faradaic effi- Reprinted by permission from Springer Nature Customer Service Centre GmbH:
ciencies in PR-EC. Additionally, H2(g) bubbles sorbed to the elec- Springer Journal of Solid State Electrochemistry, Electrochemical dissolution of alu-
trode surface could be re-oxidized to H2 O upon polarity reversal, minium in electrocoagulation experiments, Fekete et al., Copyright 2016.
which is also thermodynamically favorable over Fe(0)→Fe(II) oxi-
dation. However, it remains unclear if kinetics allow this reaction
to proceed to a significant extent. Interestingly, it has been shown treatment performance. This outcome has been particularly evi-
that terminating the current for 6 s in between each reversal with dent in Al-EC, whereas mixed results have been reported for Fe-EC.
iron electrodes decreases the buildup of the electrode surface layer Therefore, it is apparent that PRTs associated with long term treat-
by allowing anodically dissolved metal to transfer to the bulk so- ment must be optimized to avoid the adverse effects related high-
lution (Xin et al., 2018). Consequently, less iron metal is available frequency polarity reversal, which has been acknowledged before
to precipitate on the anode when it becomes the cathode. (van Genuchten et al., 2016). Furthermore, when PR-EC has been
In Al-EC, the electrochemical properties governing Al(0)→Al(III) used in field operations, polarity is reversed in between each batch
dissolution are different than those in Fe-EC. In particular, Al(III) (Amrose et al., 2014). However, it is possible that this is still too
cannot be electroplated following polarity reversal due to its frequent to avoid the detrimental effects of re-polarizing the elec-
very low reduction potential compared to the hydrolysis of wa- trodes.
ter (−1.66 V vs. −0.83 V vs. SHE [at pH 14]). The likelihood of If PR-EC is more effective at mitigating fouling and passivation
side reactions induced by polarity reversal is smaller or nonexis- in Al-EC than in Fe-EC, it could change the cost-benefit analysis
tent because aluminum does not exhibit the same reactivity as iron of EC treatment for some applications. However, as mentioned in
within the electrochemical stability range of water (Roberge, 2008). Sections 2.3.3. and 2.4., it has been reported that material precip-
In contrast, due to the anodic activation of both electrodes, the itation on the electrodes can be more significant and create oper-
protective Al2 O3 surface layer can be removed and expose more ational problems for Al-EC compared to Fe-EC for operating DC-EC
Al(0) to chemical corrosion that increases coagulant production. on the pilot scale, which may have limited the industrial applica-
For example, it has been observed that the consumption of alu- tion of Al-EC (Schulz et al., 2009). Of course, the electrode selection
minum electrodes in PR-EC is greater than in DC-EC (Pi et al., 2014; will also depend on the solution composition. For example, for re-
Wellner et al., 2018). However, similar to iron, higher Al(0) dissolu- moving some contaminants, such as Cr(VI), Fe-EC is more suitable
tion rates have been observed with longer PRTs relative to shorter (Mouedhen et al., 2009).
PRTs (Fig. 14) (Fekete et al., 2016; Hasani et al., 2019). An expla-
nation for this could be that longer PRTs allow the pH boundary 3.4. Ultrasonication
layers at the electrodes to equilibrate, which accelerates chemical
corrosion after the electrode surface layers have been removed. In It has been reported that coupling ultrasonication with EC,
contrast, there is not sufficient time to reach the required pH to referred to as sono-EC, can be an effective depassivation strat-
corrode aluminum when operating with shorter PRTs (Fig. 9). Ca- egy (He et al., 2016), while improving coagulant formation and
pacitive charging may also affect Al-EC performance at PRTs < 1 s. enhancing EC performance (Kovatcheva and Parlapanski, 1999;
A group of Soviet researchers in the 1980s took an alterna- Raschitor et al., 2014). Sonolysis can also produce radicals
tive approach, i.e., switching polarity for 5–10 s every 5–6 hrs (Eq. (35)), which mineralize organics while achieving disinfection
when operating with an aluminum anode and a titanium cathode (Wang et al., 2009). In sonolysis, the ultrasound forms microscopic
(Nikolaev et al., 1982). The purpose of this short period of polarity bubbles, referred to as cavities, that collapse once a critical radius
inversion was to remove the calcium and magnesium precipitates is attained (De Lima Leite et al., 2003). This phenomenon generates
at the cathode, whereas it was claimed that polarity reversal was points of high temperatures and pressures, which is conducive to
not effective at removing the metal oxide layer at the anode. A the formation of radicals.
built-in automated scraper was used to remove the anodic surface sonolysis
H2 O → H· + HO· (35)
layer mechanically. By applying this technique, it was reported that
the current density increased by 6–10 times (when operating po- Ultrasound also dislodges H2(g) bubbles sorbed to the elec-
tentiostatically) during long term treatment at a municipal water trodes that reduce conductivity and increase the current density
supply plant. (Kovatcheva and Parlapanski, 1999). As discussed in Section 3.3.4.,
PR-EC is seemingly effective at removing, delaying, or avoiding the presence of hydrogen on the electrodes may also trigger side
the formation of electrode surface layers, which in turn enhances reactions in PR-EC. Additionally, Sono-EC can enhance the forma-

19
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Table 5
Reported results from studies investigating sono-EC.

Electrode
Material Sample Type Current Sonication Settings Energy Consumption Conclusions Author

Fe, Al Tanning Waste 20, 40 mA 40 kHz, 350 W 435 kWh m−3 (86% Cont. sono-EC not (Maha Lakshmi and
cm−2 COD removal) recommended Sivashanmugam, 2013)
Al Laundry waste 1–7 V 43 kWz, 80 W 140 kWh m−3 (~60% Sono improved COD (Wang et al., 2009)
COD removal; removal
sonication power not
included)
Al Dye in Wastewater 2–18 mA cm−2 0–150 W 6.38 kWh m−3 (82% Sono-EC reduces (He et al., 2016)
removal) passivation
Al PO4 3− in NaHCO3 60 V 0–450W 0.31 kWh m−3 (83% 100 W: least (He et al., 2018)
Water removal) passivation, best
performance
Al Synthetic E. coli 0.5–2.5 mA 0.28 & 37 kHz; 0.212 $US m−3 Sono-EC reduced (Hashim et al., 2020)
solution, 5 g L−1 cm−2 Intermittent (5, (including treatment time by 56%
NaCl, pH 7 10 min) electricity + materials)

tion of cracks and defects in the electrode (Kovatcheva and Parla- Using a steel brush to mechanically clean iron electrodes once
panski, 1999), which can accelerate the rate of chemical corrosion a week during long-term operation reduced energy consump-
by exposing more electrode surface area. This effect could be par- tion by approximately 33% while enhancing the efficiency of ar-
ticularly beneficial in Al-EC. senic removal (Amrose et al., 2014). Many have suggested that
Continuous and intermittent sonication have both shown to be this is the most effective depassivation strategy (Bandaru et al.,
effective at impeding the onset of electrode passivation, which was 2020; Lakshmanan et al., 2009; Müller et al., 2019). However, this
confirmed using EIS (He et al., 2016). However, even though con- method requires terminating the treatment process and disassem-
tinuous sonication was more effective, coagulant formation was in- bling the EC cell, which adds labor costs. In-situ mechanical clean-
hibited, which resulted in decreased contaminant removal efficien- ing may be a viable solution, i.e., installing scrapers that remove
cies. Therefore, intermittent sonication proved to be favorable by the anodic surface layer, which has been shown to work with a
accomplishing effective depassivation while avoiding the destruc- radial electrode configuration (Nikolaev et al., 1982).
tion of flocs. Adverse effects on EC performance associated with It has also been shown that increasing the charge loading rate,
excessive sonication have been seen elsewhere (He et al., 2018). In i.e., the current density, inhibits the buildup of the electrode sur-
this study, the authors recommended operating at a lower soni- face layer over long-term Fe-EC operation (Müller et al., 2019). It
cation power to mitigate floc destruction. Another study demon- was proposed that this effect was caused by the rapid depletion of
strated that removing COD from oil tanning waste using contin- dissolved oxygen in the anodic boundary layer by an abundance of
uous sono-EC was not economically sustainable, reporting a to- electrochemically generated Fe(II). Consequently, Fe(II)→Fe(III) oxi-
tal energy consumption of 435 kWh m−3 (Maha Lakshmi and dation was depressed, which decreased the precipitation of insu-
Sivashanmugam, 2013). However, it was concluded that sono-EC lating Fe(III)-oxyhydr(oxides) on the anode surface. Operating at
was an effective depassivating strategy. Therefore, it was suggested high current densities also increases the positive charge of the
that sonication could be used when re-starting a reactor with pas- anode (Eq. (12)), which intensifies the repulsive effect between
sivated electrodes. An overview of the studies that have investi- the cationic Fe(II/III) ions and the anode (Alkhatib et al., 2020).
gated sono-EC is provided in Table 5. However, an excessively high supporting electrolyte concentration
Even though sono-EC has proven to mitigate passivation suc- will diminish the magnitude of coagulant mass transfer by elec-
cessfully while enhancing treatment performance, installing the trophoretic migration. Care should also be taken when increasing
required infrastructure, and accounting for the ensuing operating the current density due to the possibility of attracting oxyanions
costs do not make sono-EC an attractive strategy. to the anode surface, causing passivation and triggering side reac-
tions (Hu et al., 2003; Huang et al., 2009; Izquierdo et al., 2010;
Müller et al., 2019; van Genuchten et al., 2017).
3.5. Other passivation mitigation strategies Some researchers have tried operating with a horizontal elec-
trode configuration, where the cathode is placed at the bottom,
Other strategies that have been proposed for avoiding passiva- which allowed H2(g) to percolate through the solution (Abdel-
tion, fouling, and enhancing EC performance include: mechanically Aziz et al., 2020). It was claimed that this reactor design enhanced
cleaning the electrodes (Bandaru et al., 2020; Lakshmanan et al., mass transfer and decreased anodic passivation. Furthermore, us-
2009; van Genuchten et al., 2016); increasing the flow velocity to ing an oscillating anode has been shown to reduce the cell voltage
enhance hydrodynamic scouring (Chen, 2004; Ghanizadeh et al., in Fe-EC and Al-EC (Panikulam et al., 2018), which was attributed
2016; Timmes et al., 2010); and chemically cleaning the electrodes to improved mass transport of metal-oxyhydr(oxides) from the
with organic acids (Müller et al., 2019). Additionally, temporally anode. Consequently, this technique may inhibit material precip-
varying the magnitude of a static voltage with 30 min intervals itation on the anode. A “self-assembled” electrode consisting of
has shown to decrease energy consumption by 48% in continuous magnetic iron particles on a non-magnetic substrate has been de-
Al-EC and Fe-EC operations (Genc and Bakirci, 2015). Also, pulsing scribed in a patent as an effective de-fouling method (Xia et al.,
a direct current, i.e., intermittently terminating/restarting the cur- 2013). However, quantitative evidence for the performance of this
rent without reversing electrode polarity, has been demonstrated method has not been presented in the research literature.
to reduce energy consumption and sludge production by 89% and Lastly, changing the storage conditions of the electrodes in be-
30%, respectively (relative to DC-EC) (Zhou et al., 2020). This ef- tween operating cycles has been investigated by some authors
fect was attributed to the pulsed current inhibiting the formation (Müller et al., 2019; Wellner et al., 2018). It was found that stor-
a passivating surface layer. ing iron electrodes in synthetic groundwater in between cycles

20
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Table 6
Overview of the advantages and disadvantages of the different depassivation and EC performance strategies.

Strategy Advantages Disadvantages

Aggressive ion addition

• Low cost • Increases sludge production


• Simple • Risk of byproduct formation
• Effective depassivation • May require additional treatment

AC-EC

• Automatable • Low faradaic efficiency


• Uniform electrode corrosion • Uncertain effectiveness
• Low sludge production • Reduces treatment performance
• Low energy consumption by possible side reactions and
• Does not requires chemicals capacitive charging

PR-EC

• Automatable • Optimal PRT dependent on


• Uniform electrode corrosion operating conditions
• Effective depassivation • Reduces treatment performance
• Low sludge production by possible side reactions and
• Potentially more advantages for capacitive charging
Al-EC • Requires optimization
• Does not require chemicals • Requires additional electrical
infrastructure

Sono-EC

• Effective depassivation • Costly


• Could enhance EC performance • Requires additional electrical
• Does not require chemicals infrastructure

Mech. cleaning

• Effective depassivation • Requires manual labor


• Does not require additional • Requires terminating treatment
infrastructure or chemicals

Chem. cleaning

• Possibly effective depassivation • Costly


• In-situ cleaning • Requires chemical addition
• Does not require additional
infrastructure

Increased flowrate

• Possibly effective depassivation • Requires continuous pump


• In-situ cleaning • Costly
• Does not require chemicals • Does not permit batch EC

had a detrimental effect on faradaic efficiency and EC performance All these strategies have been shown to be successful at
once the reactor was restarted due to the formation of a passivat- mitigating passivation under suitable conditions. However, each
ing and coagulant-trapping surface layer (Müller et al., 2019). Oth- method has drawbacks that need to be considered. For exam-
ers found that storing aluminum electrodes in a salt solution (1 g ple, excessive aggressive ion concentrations may generate haz-
L−1 NaCl) minimized the formation of a passivating surface layer, ardous byproducts and increase sludge production. While imple-
and increased treatment performance upon re-starting the reactor menting some form of alternating electrode polarity is the most
(Wellner et al., 2018). commonly used passivation control strategy, studies have shown
that there are also drawbacks to this technique. For example, AC-
4. Conclusions EC or high-frequency PR-EC with iron electrodes may accelerate
electrode fouling while triggering side reactions that divert the
The different strategies that have been proposed to de- current and depress faradaic efficiency. PR-EC with aluminum elec-
crease/avoid passivation, and to enhance long-term EC perfor- trodes may mitigate severe passivation that can occur in some sys-
mance include: tems, which could make Al-EC a more attractive option than Fe-EC
for continuous long-term operation, depending on operating con-
• aggressive ion addition ditions. It is also more suitable to operate PR-EC with aluminum
• alternating current operation electrodes because this system is not prone to reversible electro-
• polarity reversal operation chemical reactions, which likely reduce the faradaic efficiency in
• ultrasonication Fe-EC. However, care should be taken to avoid residual aluminum
• mechanical cleaning of electrodes due to its toxicity (Singh et al., 2017). Ultrasonication is effective
• chemical cleaning of electrodes but likely too costly for the conditions investigated to date. How-
• hydrodynamic scouring

21
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

ever, this strategy could potentially be used as an in-situ electrode Asaithambi, P., Sajjadi, B., Abdul Aziz, A.R., Wan, Daud, Bin, W.M.A., 2016. Perfor-
cleaning technique in between treatment cycles. Mechanical clean- mance evaluation of hybrid electrocoagulation process parameters for the treat-
ment of distillery industrial effluent. Process Saf. Environ. Prot. 104, 406–412.
ing is seemingly the most effective strategy at decreasing passiva- doi:10.1016/j.psep.2016.09.023.
tion and enhancing EC operation. However, this strategy increases Ashraf, S.N., Rajapakse, J., Dawes, L.A., Millar, G.J., 2019. Electrocoagulation for the
labor costs and requires shutting down treatment. Therefore, it purification of highly concentrated brine produced from reverse osmosis desali-
nation of coal seam gas associated water. J. Water Process Eng. 28, 300–310.
should be avoided as much as possible. In-situ cleaning using auto- doi:10.1016/j.jwpe.2019.02.003.
mated electrode scraping mechanisms merits further investigation. Badawy, W.A., Al-Kharafi, F.M., El-Azab, A.S., 1999. Electrochemical behaviour and
A compilation of the advantages and disadvantages of each method corrosion inhibition of Al, Al-6061 and Al-Cu in neutral aqueous solutions. Cor-
ros. Sci. 41, 709–727. doi:10.1016/S0010-938X(98)00145-0.
is provided in Table 6.
Bandaru, S.R.S., Roy, A., Gadgil, A.J., van Genuchten, C.M., 2020. Long-term elec-
Based on the findings in this study, the most promising pas- trode behavior during treatment of arsenic contaminated groundwater by a
sivation control strategies are aggressive ion addition, PR-EC, and pilot-scale iron electrocoagulation system. Water Res 175, 115668. doi:10.1016/
j.watres.2020.115668.
mechanical cleaning of the electrodes. EC developers and opera-
Bao, J., Yu, W.J., Liu, Y., Wang, X., Liu, Z.Q., Duan, Y.F., 2020. Removal of per-
tors should incorporate a combination of these strategies in their fluoroalkanesulfonic acids (PFSAs) from synthetic and natural groundwater by
process designs. However, parameters such as the concentration of electrocoagulation. Chemosphere 248, 125951. doi:10.1016/j.chemosphere.2020.
aggressive ions and PRT should be optimized based on specific op- 125951.
Bard, A.J., Jordan, J., Parsons, R., 1985. Standard Potentials in Aqueous Solution. Mar-
erating conditions. Failing to optimize a strategy could lead to de- cel Dekker, 1st ed.
creased treatment performance or operational challenges. There- Ben Grich, N., Attour, A., Le Page Mostefa, M., Guesmi, S., Tlili, M., Lapicque, F., 2019.
fore, the processes, passivation mechanisms, and depassivation Fluoride removal from water by electrocoagulation: Effect of the type of water
and the experimental parameters. Electrochim. Acta 316, 257–265. doi:10.1016/
strategies discussed in this article should be considered when as- j.electacta.2019.05.130.
signing operating parameters in the design of EC systems. The se- Betancor-Abreu, A., Mena, V.F., González, S., Delgado, S., Souto, R.M., Santana, J.J.,
lection of electrode material should also be carefully considered 2019. Design and optimization of an electrocoagulation reactor for fluoride re-
mediation in underground water sources for human consumption. J. Water Pro-
because passivation characteristics can vary significantly depend- cess Eng. 31, 100865. doi:10.1016/j.jwpe.2019.100865.
ing on the application. Bian, Y., Ge, Z., Albano, C., Lobo, F.L., Ren, Z.J., 2019. Oily bilge water treatment using
It is also clear that the nature and mechanisms of depassivation DC/AC powered electrocoagulation. Environ. Sci. Water Res. Technol. 5, 1654–
1660. doi:10.1039/c9ew00497a.
by aggressive ion addition and polarity reversal are sensitive to the
Cañizares, P., Carmona, M., Lobato, J., Martínez, F., Rodrigo, M.A., 2005. Electrodisso-
water composition, and are not well understood. Further studies of lution of aluminum electrodes in electrocoagulation processes. Ind. Eng. Chem.
the mechanisms of passivation and depassivation as a function of Res. 44, 4178–4185. doi:10.1021/ie048858a.
Cañizares, P., Jiménez, C., Martínez, F., Sáez, C., Rodrigo, M.A., 2007. Study of the
water composition and operating conditions are needed to provide
electrocoagulation process using aluminum and iron electrodes. Ind. Eng. Chem.
deeper insight into these processes. Lastly, studies investigating the Res. 46, 6189–6195. doi:10.1021/ie070059f.
effectiveness of implementing depassivation strategies during long- Cerqueira, A.A., Souza, P.S.A., Marques, M.R.C., 2014. Effects of direct and alternating
term continuous EC operation are needed. More research is also current on the treatment of oily water in an electroflocculation process. Brazil-
ian J. Chem. Eng. 31, 693–701. doi:10.1590/0104-6632.20140313s0 0 0 02363.
required to further understand the relationship between EC perfor- Cesar Lopes Geraldino, H., Izabelle Simionato, J., Karoliny Formicoli de Souza Fre-
mance, solution composition, and operating conditions. itas, T., Carla Garcia, J., Evelázio de Souza, N., 2016. Evaluation of the electrode
wear and the residual concentration of iron in a system of electrocoagulation.
Desalin. Water Treat. 57, 13377–13387. doi:10.1080/19443994.2015.1058192.
Funding Chellam, S., Sari, M.A., 2016. Aluminum electrocoagulation as pretreatment during
microfiltration of surface water containing NOM: A review of fouling, NOM, DBP,
and virus control. J. Hazard. Mater. 304, 490–501. doi:10.1016/j.jhazmat.2015.10.
This research was funded by the Natural Science and Engineer- 054.
ing Research Council of Canada (STPGP 506,951–2017). Chen, G., 2004. Electrochemical technologies in wastewater treatment. Sep. Purif.
Technol. 38, 11–41. doi:10.1016/j.seppur.2003.10.006.
Chen, M., Dollar, O., Shafer-Peltier, K., Randtke, S., Waseem, S., Peltier, E., 2020.
Boron removal by electrocoagulation: Removal mechanism, adsorption models
Declaration of Competing Interest
and factors influencing removal. Water Res 170, 115362. doi:10.1016/j.watres.
2019.115362.
The authors declare that they have no known competing finan- Chen, X., Chen, G., Yue, P.L., 2002. Investigation on the electrolysis voltage of
cial interests or personal relationships that could have appeared to electrocoagulation. Chem. Eng. Sci. 57, 2449–2455. doi:10.1016/S0 0 09-2509(02)
00147-1.
influence the work reported in this paper. Chen, X., Chen, G., Yue, P.L., 20 0 0. Separation of pollutants from restaurant
wastewater by electrocoagulation. Sep. Purif. Technol. 19, 65–76. doi:10.1016/
S1383-5866(99)0 0 072-6.
Acknowledgments Chow, H., Pham, A.L.T., 2019. Effective removal of silica and sulfide from oil sands
thermal in-situ produced water by electrocoagulation. J. Hazard. Mater. 380,
120880. doi:10.1016/j.jhazmat.2019.120880.
The authors would like to thank Dr. Anh Pham and Héline Chow Conway, B.E., 2018. Electrochemical Double-Layer Capacitors. In: Fuller, T.F.,
at the University of Waterloo for sharing their expertise during the Harb, J.N. (Eds.), Electrochemical Engineering. 1st ed. Wiley, New York,
writing process. pp. 251–275.
Cook, D.C., 2005. Spectroscopic identification of protective and non-protective corro-
sion coatings on steel structures in marine environments. Corros. Sci. 47, 2550–
References 2570. doi:10.1016/j.corsci.2004.10.018.
Daneshvar, N., Khataee, A.R., Amani Ghadim, A.R., Rasoulifard, M.H., 2007. Decol-
Abdel-Aziz, M.H., El-Ashtoukhy, E.S.Z., Sh. Zoromba, M., Bassyouni, M., orization of C.I. Acid Yellow 23 solution by electrocoagulation process: Inves-
Sedahmed, G.H., 2020. Removal of nitrates from water by electrocoagula- tigation of operational parameters and evaluation of specific electrical energy
tion using a cell with horizontally oriented Al serpentine tube anode. J. Ind. consumption (SEEC). J. Hazard. Mater. 148, 566–572. doi:10.1016/j.jhazmat.2007.
Eng. Chem. 82, 105–112. doi:10.1016/j.jiec.2019.10.001. 03.028.
Alkhatib, A.M., Hawari, A.H., Hafiz, M.A., Benamor, A., 2020. A novel cylindrical elec- De Lima Leite, R.H., Cognet, P., Wilhelm, A.M., Delmas, H., 2003. Anodic oxidation of
trode configuration for inducing dielectrophoretic forces during electrocoagula- 2,4-dihydroxybenzoic acid for wastewater treatment. J. Appl. Electrochem. 33,
tion. J. Water Process Eng. 35, 101195. doi:10.1016/j.jwpe.2020.101195. 693–701. doi:10.1023/A:1025056001368.
Amrose, S.E., Bandaru, S.R.S., Delaire, C., van Genuchten, C.M., Dutta, A., Deb- Dermentzis, K., Christoforidis, A., Valsamidou, E., Lazaridou, A., Kokkinos, N., 2011.
Sarkar, A., Orr, C., Roy, J., Das, A., Gadgil, A.J., 2014. Electro-chemical arsenic Removal of hexavalent chromium from electroplating wastewater by electro-
remediation: Field trials in West Bengal. Sci. Total Environ. 488–489, 539–546. coagulation with iron electrodes. Glob. Nest J. 13, 412–418. doi:10.30955/gnj.
doi:10.1016/j.scitotenv.2013.11.074. 0 0 0770.
Arroyo, M.G., Pérez-Herranz, V., Montañés, M.T., García-Antón, J., Guiñón, J.L., 2009. Doggaz, A., Attoura, A., Le Page Mostefa, M., Côme, K., Tlili, M., Lapicque, F.,
Effect of pH and chloride concentration on the removal of hexavalent chromium 2019. Removal of heavy metals by electrocoagulation from hydrogenocarbonate-
in a batch electrocoagulation reactor. J. Hazard. Mater. 169, 1127–1133. doi:10. containing waters: Compared cases of divalent iron and zinc cations. J. Water
1016/j.jhazmat.2009.04.089. Process Eng. 29, 100796. doi:10.1016/j.jwpe.2019.100796.

22
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Donaldson, J.D., Grimes, S.M., Yasri, N.G., Wheals, B., Parrick, J., Errington, W.E., Gorski, C.A., Edwards, R., Sander, M., Hofstetter, T.B., Stewart, S.M., 2016. Thermo-
2002. Anodic oxidation of the dye materials methylene blue, acid blue 25, re- dynamic Characterization of Iron Oxide-Aqueous Fe2+ Redox Couples. Environ.
active blue 2 and reactive blue 15 and the characterisation of novel interme- Sci. Technol. 50, 8538–8547. doi:10.1021/acs.est.6b02661.
diate compounds in the anodic oxidation of methylene blue. J. Chem. Technol. Gotsi, M., Kalogerakis, N., Psillakis, E., Samaras, P., Mantzavinos, D., 2005. Elec-
Biotechnol. 77, 756–760. doi:10.1002/jctb.642. trochemical oxidation of olive oil mill wastewaters. Water Res 39, 4177–4187.
Donneys-Victoria, D., Marriaga-Cabrales, N., Machuca-Martínez, F., Benavides- doi:10.1016/j.watres.2005.07.037.
Guerrero, J., Cloutier, S.G., 2020. Indigo carmine and chloride ions removal by Grøterud, O., Smoczynski, L., 1986. Phosphorus removal from water by means of
electrocoagulation. Simultaneous production of brucite and layered double hy- electrolysis. Water Res 20, 667–669. doi:10.1016/0 043-1354(86)90 032-1.
droxides. J. Water Process Eng. 33, 101106. doi:10.1016/j.jwpe.2019.101106. Guan, X., Enalls, B.C., Clarke, D.R., Girguis, P., 2017. Iron sulfide formation on iron
Dubrawski, K.L., Mohseni, M., 2013. In-situ identification of iron electrocoagulation substrates by electrochemical reaction in anoxic conditions. Cryst. Growth Des.
speciation and application for natural organic matter (NOM) removal. Water Res 17, 6332–6340. doi:10.1021/acs.cgd.7b01013.
47, 5371–5380. doi:10.1016/j.watres.2013.06.021. Hasani, G., Maleki, A., Daraei, H., Ghanbari, R., Safari, M., McKay, G., Yetilmez-
Dubrawski, K.L., Van Genuchten, C.M., Delaire, C., Amrose, S.E., Gadgil, A.J., soy, K., Ilhan, F., Marzban, N., 2019. A comparative optimization and perfor-
Mohseni, M., 2015. Production and transformation of mixed-valent nanoparti- mance analysis of four different electrocoagulation-flotation processes for hu-
cles generated by Fe(0) electrocoagulation. Environ. Sci. Technol. 49, 2171–2179. mic acid removal from aqueous solutions. Process Saf. Environ. Prot. 121, 103–
doi:10.1021/es505059d. 117. doi:10.1016/j.psep.2018.10.025.
Dura, A., Breslin, C.B., 2019. The removal of phosphates using electrocoagulation Hashim, K.S., Ali, S.S.M., AlRifaie, J.K., Kot, P., Shaw, A., Al Khaddar, R., Id-
with Al−Mg anodes. J. Electroanal. Chem. 846, 113161. doi:10.1016/j.jelechem. owu, I., Gkantou, M., 2020. Escherichia coli inactivation using a hybrid
2019.05.043. ultrasonic–electrocoagulation reactor. Chemosphere 247, 125868. doi:10.1016/j.
El-Ghenymy, A., Alsheyab, M., Khodary, A., Sirés, I., Abdel-Wahab, A., 2020. Corrosion chemosphere.2020.125868.
behavior of pure titanium anodes in saline medium and their performance for He, C.C., Hu, C.Y., Lo, S.L., 2018. Integrating chloride addition and ultrasonic pro-
humic acid removal by electrocoagulation. Chemosphere 246, 125674. doi:10. cessing with electrocoagulation to remove passivation layers and enhance phos-
1016/j.chemosphere.2019.125674. phate removal. Sep. Purif. Technol. 201, 148–155. doi:10.1016/j.seppur.2018.03.
Emamjomeh, M.M., Sivakumar, M., 2009. Review of pollutants removed by electro- 011.
coagulation and electrocoagulation/flotation processes. J. Environ. Manage. 90, He, C.C., Hu, C.Y., Lo, S.L., 2016. Evaluation of sono-electrocoagulation for the re-
1663–1679. doi:10.1016/j.jenvman.2008.12.011. moval of Reactive Blue 19 passive film removed by ultrasound. Sep. Purif. Tech-
Esmailzadeh, S., Aliofkhazraei, M., Sarlak, H., 2018. Interpretation of Cyclic Po- nol. 165, 107–113. doi:10.1016/j.seppur.2016.03.047.
tentiodynamic Polarization Test Results for Study of Corrosion Behavior of Heidmann, I., Calmano, W., 2008. Removal of Cr(VI) from model wastewaters by
Metals: A Review. Prot. Met. Phys. Chem. Surfaces 54, 976–989. doi:10.1134/ electrocoagulation with Fe electrodes. Sep. Purif. Technol. 61, 15–21. doi:10.1016/
S207020511805026X. j.seppur.2007.09.011.
Eyvaz, M., 2016. Treatment of brewery wastewater with electrocoagulation: Improv- Hernández-Espejel, A., Palomar-Pardavé, M., Cabrera-Sierra, R., Romero-Romo, M.,
ing the process performance by using alternating pulse current. Int. J. Elec- Ramírez-Silva, M.T., Arce-Estrada, E.M., 2011. Kinetics and mechanism of the
trochem. Sci. 11, 4988–5008. doi:10.20964/2016.06.11. electrochemical formation of iron oxidation products on steel immersed in sour
Eyvaz, M., Kirlaroglu, M., Aktas, T.S., Yuksel, E., 2009. The effects of alternating cur- acid media. J. Phys. Chem. B 115, 1833–1841. doi:10.1021/jp106851b.
rent electrocoagulation on dye removal from aqueous solutions. Chem. Eng. J. Holt, P.K., Barton, G.W., Mitchell, C.A., 2005. The future for electrocoagulation as a
153, 16–22. doi:10.1016/j.cej.2009.05.028. localised water treatment technology. Chemosphere 59, 355–367. doi:10.1016/j.
Fayad, N., Yehya, T., Audonnet, F., Vial, C., 2017. Preliminary purification of volatile chemosphere.2004.10.023.
fatty acids in a digestate from acidogenic fermentation by electrocoagulation. Holt, P.K., Barton, G.W., Wark, M., Mitchell, C.A., 2002. A quantitative comparison
Sep. Purif. Technol. 184, 220–230. doi:10.1016/j.seppur.2017.04.041. between chemical dosing and electrocoagulation. Colloids Surfaces A Physic-
Fekete, É., Lengyel, B., Cserfalvi, T., Pajkossy, T., 2016. Electrochemical dissolution ochem. Eng. Asp. 211, 233–248. doi:10.1016/S0927- 7757(02)00285- 6.
of aluminium in electrocoagulation experiments. J. Solid State Electrochem. 20, Hruska, L.W., Savinell, R.F., 1981. Investigation of factors affecting performance of
3107–3114. doi:10.10 07/s10 0 08- 016- 3195- 6. the iron-redox battery. J. Electrochem. Soc. 128, 18–25. doi:10.1149/1.2127366.
Ferro, S., De Battisti, A., Duo, I., Comninellis, C., Haenni, W., Perret, A., 20 0 0. Chlorine Hu, C.Y., Lo, S.L., Kuan, W.H., 2003. Effects of co-existing anions on fluoride removal
Evolution at Highly Boron-Doped Diamond Electrodes. J. Electrochem. Soc. 147, in electrocoagulation (EC) process using aluminum electrodes. Water Res 37,
2614. doi:10.1149/1.1393578. 4513–4523. doi:10.1016/S0 043-1354(03)0 0378-6.
Fuladpanjeh-Hojaghan, B., Elsutohy, M.M., Kabanov, V., Heyne, B., Trifkovic, M., Huang, C.H., Chen, L., Yang, C.L., 2009. Effect of anions on electrochemical coagula-
Roberts, E.P.L., 2019. In-Operando Mapping of pH Distribution in Electro- tion for cadmium removal. Sep. Purif. Technol. 65, 137–146. doi:10.1016/j.seppur.
chemical Processes. Angew. Chem. Int. Ed. 58, 16815–16819. doi:10.1002/anie. 2008.10.029.
201909238. Izquierdo, C.J., Canizares, P., Rodrigo, M.A., Leclerc, J.P., Valentin, G., Lapicque, F.,
Gao, S., Du, M., Tian, J., Yang, Jianyu, Yang, Jixian, Ma, F., Nan, J., 2010. Effects of 2010. Effect of the nature of the supporting electrolyte on the treatment of solu-
chloride ions on electro-coagulation-flotation process with aluminum electrodes ble oils by electrocoagulation. Desalination 255, 15–20. doi:10.1016/j.desal.2010.
for algae removal. J. Hazard. Mater. 182, 827–834. doi:10.1016/j.jhazmat.2010.06. 01.022.
114. Jiang, J., Graham, N., Andre, C., Kelsall, G.H., Brandon, N., 2002. Laboratory study
Garcia-Segura, S., Eiband, M.M.S.G., de Melo, J.V., Martínez-Huitle, C.A., 2017. Elec- of electro-coagulation-flotation for water treatment. Water Res. 36, 4064–4078.
trocoagulation and advanced electrocoagulation processes: A general review doi:10.1016/S0 043-1354(02)0 0118-5.
about the fundamentals, emerging applications and its association with other Jianqian, J., 1988. An Anodic Passivation of Electrocoagulator in the Process of Water
technologies. J. Electroanal. Chem. 801, 267–299. doi:10.1016/j.jelechem.2017.07. Treatment. Water Treat 3, 344–352.
047. Kabdaşli, I., Vardar, B., Arslan-Alaton, I., Tünay, O., 2009. Effect of dye auxiliaries on
Genc, A., Bakirci, B., 2015. Treatment of emulsified oils by electrocoagulation: Pulsed color and COD removal from simulated reactive dyebath effluent by electroco-
voltage applications. Water Sci. Technol. 71, 1196–1202. doi:10.2166/wst.2015. agulation. Chem. Eng. J. 148, 89–96. doi:10.1016/j.cej.20 08.08.0 06.
092. Kabdaşlı, I., Arslan-Alaton, I., Ölmez-Hancı, T., Tünay, O., 2012. Electrocoagulation
Gerónimo-López, C., Vazquez-Arenas, J., Picquart, M., González, I., 2014. The ener- applications for industrial wastewaters: a critical review. Environ. Technol. Rev.
getic conditions determining the active dissolution of carbon steel during elec- 1, 2–45. doi:10.1080/21622515.2012.715390.
trocoagulation in sulfate media. Electrochim. Acta 136, 146–156. doi:10.1016/j. Kamaraj, R., Ganesan, P., Lakshmi, J., Vasudevan, S., 2013. Removal of copper
electacta.2014.05.069. from water by electrocoagulation process-effect of alternating current (AC)
Ghanizadeh, G., Shariati neghab, G., Salem, M., Khalagi, K., 2016. Taguchi experi- and direct current (DC). Environ. Sci. Pollut. Res. 20, 399–412. doi:10.1007/
mental design for electrocoagulation process using alternating and direct cur- s11356- 012- 0855- 7.
rent on fluoride removal from water. Desalin. Water Treat. 57, 12675–12683. Karamati-Niaragh, E., Alavi Moghaddam, M.R., Emamjomeh, M.M.„ Nazlabadi, E.,
doi:10.1080/19443994.2015.1049562. 2019. Evaluation of direct and alternating current on nitrate removal us-
Ghernaout, D., 2013. Advanced oxidation phenomena in electrocoagulation process: ing a continuous electrocoagulation process: Economical and environmental
a myth or a reality? Desalin. Water Treat 51, 7536–7554. doi:10.1080/19443994. approaches through RSM. J. Environ. Manage. 230, 245–254. doi:10.1016/j.
2013.792520. jenvman.2018.09.091.
Ghernaout, D., Ghernaout, B., 2011. On the controversial effect of sodium sulphate as Keshmirizadeh, E., Yousefi, S., Rofouei, M.K., 2011. An investigation on the new
supporting electrolyte on electrocoagulation process: A review. Desalin. Water operational parameter effective in Cr(VI) removal efficiency: A study on elec-
Treat. 27, 243–254. doi:10.5004/dwt.2011.1983. trocoagulation by alternating pulse current. J. Hazard. Mater. 190, 119–124.
Gobbi, L.C.A., Nascimento, I.L., Muniz, E.P., Rocha, S.M.S., Porto, P.S.S., 2018. Electro- doi:10.1016/j.jhazmat.2011.03.010.
coagulation with polarity switch for fast oil removal from oil in water emul- Khandegar, V., Saroha, A.K., 2013. Electrocoagulation for the treatment of textile
sions. J. Environ. Manage. 213, 119–125. doi:10.1016/j.jenvman.2018.01.069. industry effluent - A review. J. Environ. Manage. 128, 949–963. doi:10.1016/j.
Golder, A.K., Chanda, A.K., Samanta, A.N., Ray, S., 2011. Removal of hexavalent jenvman.2013.06.043.
chromium by electrochemical reduction-precipitation: Investigation of pro- Kobya, M., Hiz, H., Senturk, E., Aydiner, C., Demirbas, E., 2006. Treatment of potato
cess performance and reaction stoichiometry. Sep. Purif. Technol. 76, 345–350. chips manufacturing wastewater by electrocoagulation. Desalination 190, 201–
doi:10.1016/j.seppur.2010.11.002. 211. doi:10.1016/j.desal.2011.09.039.
Golder, A.K., Samanta, A.N., Ray, S., 2007. Removal of Cr3+ by electrocoagula- Kovatcheva, V.K., Parlapanski, M.D., 1999. Sono-electrocoagulation of iron hydrox-
tion with multiple electrodes: Bipolar and monopolar configurations. J. Hazard. ides. Colloids Surfaces A Physicochem. Eng. Asp. 149, 603–608. doi:10.1016/
Mater. 141, 653–661. doi:10.1016/j.jhazmat.2006.07.025. S0927- 7757(98)00414- 2.

23
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Lakshmanan, D., Clifford, D.A., Samanta, G., 2009. Ferrous and ferric ion gen- Müller, S., Behrends, T., van Genuchten, C.M., 2019. Sustaining efficient production of
eration during iron electrocoagulation. Environ. Sci. Technol. 43, 3853–3859. aqueous iron during repeated operation of Fe(0)-electrocoagulation. Water Res
doi:10.1021/es8036669. 155, 455–464. doi:10.1016/j.watres.2018.11.060.
Lakshmipathiraj, P., Bhaskar Raju, G., Raviatul Basariya, M., Parvathy, S., Prab- Nikolaev, N.V., Kozlovskii, A.S., Utkin, I.I., 1982. Treating Natural Waters in Small Wa-
hakar, S., 2008. Removal of Cr (VI) by electrochemical reduction. Sep. Purif. ter Systems by Filtration with Electrocoagulation. Khimiya i Tekhnologiya Vody
Technol. 60, 96–102. doi:10.1016/j.seppur.2007.07.053. 4, 244–247.
Lee, W.J., Pyun, S.Il, 20 0 0. Effects of sulphate ion additives on the pitting corrosion Ölmez, T., 2009. The optimization of Cr(VI) reduction and removal by electrocoag-
of pure aluminum in 0.01 M NaCl solution. Electrochim. Acta 45, 1901–1910. ulation using response surface methodology. J. Hazard. Mater. 162, 1371–1378.
doi:10.1016/S0013- 4686(99)00418- 1. doi:10.1016/j.jhazmat.2008.06.017.
Liu, Y., Hu, X.M., Zhao, Y., Wang, J., Lu, M.X., Peng, F.H., Bao, J., 2018. Removal Ozyonar, F., Karagozoglu, B., 2015. Treatment of pretreated coke wastewater by elec-
of perfluorooctanoic acid in simulated and natural waters with different elec- trocoagulation and electrochemical peroxidation processes. Sep. Purif. Technol.
trode materials by electrocoagulation. Chemosphere 201, 303–309. doi:10.1016/ 150, 268–277. doi:10.1016/j.seppur.2015.07.011.
j.chemosphere.2018.02.129. Pally, D., Le Bescop, P., Schlegel, M.L., Miserque, F., Chomat, L., Neff, D., L’Hostis, V.,
Lobo, F.L., Wang, H., Huggins, T., Rosenblum, J., Linden, K.G., Ren, Z.J., 2016. Low- 2020. Corrosion behavior of iron plates in cementitious solution at 80°C in
energy hydraulic fracturing wastewater treatment via AC powered electrocoag- anaerobic conditions. Corros. Sci. 170, 108650. doi:10.1016/j.corsci.2020.108650.
ulation with biochar. J. Hazard. Mater. 309, 180–184. doi:10.1016/j.jhazmat.2016. Panikulam, P.J., Yasri, N., Roberts, E.P.L., 2018. Electrocoagulation using an oscillating
02.020. anode for kaolin removal. J. Environ. Chem. Eng. 6, 2785–2793. doi:10.1016/j.
Lumsden, J.B., Stocker, P.J., Tsai, S.C., 1981. The composition and morphology of pits jece.2018.04.020.
formed on iron in an inhibited chloride solution. Appl. Surf. Sci. 7, 347–354. Phalakornkule, C., Sukkasem, P., Mutchimsattha, C., 2010. Hydrogen recovery from
doi:10.1016/0378- 5963(81)90082- 9. the electrocoagulation treatment of dye-containing wastewater. Int. J. Hydrogen
Maha Lakshmi, P., Sivashanmugam, P., 2013. Treatment of oil tanning effluent by Energy 35, 10934–10943. doi:10.1016/j.ijhydene.2010.06.100.
electrocoagulation: Influence of ultrasound and hybrid electrode on COD re- Pi, K.W., Gao, L.X., Li, Z., Wang, M., Huang, L., 2011. PAC with high content of Al13
moval. Sep. Purif. Technol. 116, 378–384. doi:10.1016/j.seppur.2013.05.026. polymer prepared by electrolysis with periodical reversal of electrodes. Colloids
Maher, E.K., O’Malley, K.N., Heffron, J., Huo, J., Mayer, B.K., Wang, Y., McNamara, P.J., Surfaces A Physicochem. Eng. Asp. 387, 113–117. doi:10.1016/j.colsurfa.2011.07.
2019. Analysis of operational parameters, reactor kinetics, and floc characteri- 040.
zation for the removal of estrogens via electrocoagulation. Chemosphere 220, Pi, K.W., Gong, W.Q., Wang, M., Huang, Z.Q., 2008. Enhancing electrochemical prepa-
1141–1149. doi:10.1016/j.chemosphere.2018.12.161. ration of polyaluminum chloride by a magnetic field. Colloids Surfaces A Physic-
Mamelkina, M.A., Cotillas, S., Lacasa, E., Sáez, C., Tuunila, R., Sillanpää, M., Häkki- ochem. Eng. Asp. 330, 103–107. doi:10.1016/j.colsurfa.2008.07.051.
nen, A., Rodrigo, M.A., 2017. Removal of sulfate from mining waters by electro- Pi, K.W., Xiao, Q., Zhang, H.Q., Xia, M., Gerson, A.R., 2014. Decolorization of syn-
coagulation. Sep. Purif. Technol. 182, 87–93. doi:10.1016/j.seppur.2017.03.044. thetic Methyl Orange wastewater by electrocoagulation with periodic reversal
Mansoorian, H.J., Mahvi, A.H., Jafari, A.J., 2014. Removal of lead and zinc from bat- of electrodes and optimization by RSM. Process Saf. Environ. Prot. 92, 796–806.
tery industry wastewater using electrocoagulation process: Influence of direct doi:10.1016/j.psep.2014.02.008.
and alternating current by using iron and stainless steel rod electrodes. Sep. Price, N., Prasad, P., Gaffel, J., 2018. Literature Review – Evaluation of Electro-
Purif. Technol. 135, 165–175. doi:10.1016/j.seppur.2014.08.012. coagulation as a wastewater treatment technology for meat processors. Aus-
Mansouri, K., Ibrik, K., Bensalah, N., Abdel-Wahab, A., 2011. Anodic dissolution of tralian Meat Processor Corporation https://www.ampc.com.au/uploads/cgblog/
pure aluminum during electrocoagulation process: Influence of supporting elec- id416/2018_1140_Milestone_3_Electroco_(1).pdf.
trolyte, initial pH, and current density. Ind. Eng. Chem. Res. 50, 13362–13372. Puigdomenech, I., 2019. Medusa-Hydra chemical equilibrium software. Downloaded
doi:10.1021/ie201206d. from https://www.kth.se/che/medusa/downloads-1.386254, February 2020.
Mao, X., Hong, S., Zhu, H., Lin, H., Wei, L., Gan, F., 2008. Alternating pulse current Pyun, S.I., Moon, S.M., Ahn, S.H., Kim, S.S., 1999. Effects of Cl-, NO3- and SO42-
in electrocoagulation for wastewater treatment to prevent the passivation of ions on anodic dissolution of pure aluminum in alkaline solution. Corros. Sci.
al electrode. J. Wuhan Univ. Technol. Mater. Sci. Ed. 23, 239–241. doi:10.1007/ 41, 653–667. doi:10.1016/S0010-938X(98)00132-2.
s11595- 006- 2239- 7. Ragush, C.M., Schmidt, J.J., Krkosek, W.H., Gagnon, G.A., Truelstrup-Hansen, L.,
Martin, F.J., Cheek, G.T., O’Grady, W.E., Natishan, P.M., 2005. Impedance studies of Jamieson, R.C., 2015. Performance of municipal waste stabilization ponds in the
the passive film on aluminium. Corros. Sci. 47, 3187–3201. doi:10.1016/j.corsci. Canadian Arctic. Ecol. Eng. 83, 413–421. doi:10.1016/j.ecoleng.2015.07.008.
2005.05.058. Raschitor, A., Fernandez, C.M., Cretescu, I., Rodrigo, M.A., Cañizares, P., 2014. Sono-
McBeath, S.T., Nouri-Khorasani, A., Mohseni, M., Wilkinson, D.P., 2020. In-situ deter- electrocoagulation of wastewater polluted with Rhodamine 6G. Sep. Purif. Tech-
mination of current density distribution and fluid modeling of an electrocoag- nol. 135, 110–116. doi:10.1016/j.seppur.2014.08.003.
ulation process and its effects on natural organic matter removal for drinking Refaey, S.A.M., 2005. Inhibition of steel pitting corrosion in HCl by some inorganic
water treatment. Water Res 171, 115404. doi:10.1016/j.watres.2019.115404. anions. Appl. Surf. Sci. 240, 396–404. doi:10.1016/j.apsusc.2004.07.014.
Mechelhoff, M., 2008. Electrochemical Electrocoagulation Investigation of Reactors Ren, M., Song, Y., Xiao, S., Zeng, P., Peng, J., 2011. Treatment of berberine hydrochlo-
for Water Purification PhD thesis. Imperial College London, UK. ride wastewater by using pulse electro-coagulation process with Fe electrode.
Mechelhoff, M., Kelsall, G.H., Graham, N.J.D., 2013a. Electrochemical behaviour of Chem. Eng. J. 169, 84–90. doi:10.1016/j.cej.2011.02.056.
aluminium in electrocoagulation processes. Chem. Eng. Sci. 95, 301–312. doi:10. Rickard, D., Luther, G.W., 2007. Chemistry of iron sulfides. Chem. Rev. 107, 514–562.
1016/j.ces.2013.03.010. doi:10.1021/cr0503658.
Mechelhoff, M., Kelsall, G.H., Graham, N.J.D., 2013b. Super-faradaic charge yields for Roberge, P.R., 2008. Corrosion Engineering: Principles and Practice, 1st ed McGraw
aluminium dissolution in neutral aqueous solutions. Chem. Eng. Sci. 95, 353– Hill, New York.
359. doi:10.1016/j.ces.2013.03.016. Sahu, O., Mazumdar, B., Chaudhari, P.K., 2014. Treatment of wastewater by elec-
Miller, J., 1977. Electrolytic cell for treatment of water. trocoagulation: A review. Environ. Sci. Pollut. Res. 21, 2397–2413. doi:10.1007/
Miller, J., 1978. Electrolytic cell for treatment of water. s11356-013-2208-6.
Mohora, E., Rončević, S., Agbaba, J., Tubić, A., Mitić, M., Klašnja, M., Dalmacija, B., Sasson, M.Ben, Calmano, W., Adin, A., 2009. Iron-oxidation processes in an elec-
2014. Removal of arsenic from groundwater rich in natural organic matter troflocculation (electrocoagulation) cell. J. Hazard. Mater. 171, 704–709. doi:10.
(NOM) by continuous electrocoagulation/flocculation (ECF). Sep. Purif. Technol. 1016/j.jhazmat.2009.06.057.
136, 150–156. doi:10.1016/j.seppur.2014.09.006. Schmuki, P., 2002. From Bacon to barriers: A review on the passivity of metals and
Mollah, M.Y.A., Morkovsky, P., Gomes, J.A.G., Kesmez, M., Parga, J., Cocke, D.L., 2004a. alloys. J. Solid State Electrochem. 6, 145–164. doi:10.1007/s100080100219.
Fundamentals, present and future perspectives of electrocoagulation. J. Hazard. Schulz, M.C., Baygents, J.C., Farrell, J., 2009. Laboratory and pilot testing of electro-
Mater. 114, 199–210. doi:10.1016/j.jhazmat.20 04.08.0 09. coagulation for removing scaleforming species from industrial process waters.
Mollah, M.Y.A., Pathak, S.R., Patil, P.K., Vayuvegula, M., Agrawal, T.S., Gomes, J.A.G., Int. J. Environ. Sci. Technol. 6, 521–526. doi:10.1007/bf03326091.
Kesmez, M., Cocke, D.L., 2004b. Treatment of orange II azo-dye by electrocoag- Secula, M.S., Cretescu, I., Cagnon, B., Manea, L.R., Stan, C.S., Breaban, I.G., 2013. Frac-
ulation (EC) technique in a continuous flow cell using sacrificial iron electrodes. tional factorial design study on the performance of GAC-enhanced electrocoag-
J. Hazard. Mater. 109, 165–171. doi:10.1016/j.jhazmat.2004.03.011. ulation process involved in color removal from dye solutions. Materials (Basel)
Mollah, M.Y.A., Schennach, R., Parga, J.R., Cocke, D.L., 2001. Electrocoagulation 6, 2723–2746. doi:10.3390/ma6072723.
(EC)- Science and applications. J. Hazard. Mater. 84, 29–41. doi:10.1016/ Shamaei, L., Khorshidi, B., Perdicakis, B., Sadrzadeh, M., 2018. Treatment of oil sands
S0304-3894(01)00176-5. produced water using combined electrocoagulation and chemical coagulation
Moreno, C., H, A., Cocke, D.L., Gromes, J.A., Morkovsky, P., Parga, J.R., Peterson, E., techniques. Sci. Total Environ. 645, 560–572. doi:10.1016/j.scitotenv.2018.06.387.
Garcia, C., 2009. Electrochemical reactions for electrocoagulation using iron Shao-qin, Y.U., Dao-wen, X., Xiang-bin, C., Jiang-jun, Y.O.U., Shan, X., He-de, W.,
electrodes. Ind. Eng. Chem. Res. 48, 2275–2282. doi:10.1021/ie8013007. Zhongnan, H., Corporation, E., 2014. Study on the treatment of chromium-con-
Mouedhen, G., Feki, M., De Petris-Wery, M., Ayedi, H.F., 2009. Electrochemical re- taining wastewater by cycle-direction electro-flocculation. China Environ. Sci.
moval of Cr(VI) from aqueous media using iron and aluminum as electrode ma- 34, 118–122.
terials: Towards a better understanding of the involved phenomena. J. Hazard. Singh, Shweta, Tripathi, D.K., Singh, Swati, Sharma, S., Dubey, N.K., Chauhan, D.K.,
Mater. 168, 983–991. doi:10.1016/j.jhazmat.2009.02.117. Vaculík, M., 2017. Toxicity of aluminium on various levels of plant cells and or-
Mouedhen, G., Feki, M., Wery, M.D.P., Ayedi, H.F., 2008. Behavior of aluminum elec- ganism: A review. Environ. Exp. Bot. 137, 177–193. doi:10.1016/j.envexpbot.2017.
trodes in electrocoagulation process. J. Hazard. Mater. 150, 124–135. doi:10. 01.005.
1016/j.jhazmat.2007.04.090.
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.
Manage. 186, 24–41. doi:10.1016/j.jenvman.2016.10.032.

24
M. Ingelsson, N. Yasri and E.P.L. Roberts Water Research 187 (2020) 116433

Smoczyński, L., Kalinowski, S., Ratnaweera, H., Kosobucka, M., Trifescu, M., Pieczulis- Wu, W., Huang, Z.H., Lim, T.T., 2014. Recent development of mixed metal oxide an-
Smoczyńska, K., 2017. Electrocoagulation of municipal wastewater – a pilot-scale odes for electrochemical oxidation of organic pollutants in water. Appl. Catal. A
test. Desalin. Water Treat. 72, 162–168. doi:10.5004/dwt.2017.20645A. Gen. 480, 58–78. doi:10.1016/j.apcata.2014.04.035.
Tartakovsky, B., Kleiner, Y., Manuel, M.F., 2018. Bioelectrochemical anaerobic sewage Xia, Z., Yang, H., Sui, C., Sun, Y., Cai, W., 2013. An Apparatus And Method For Elec-
treatment technology for Arctic communities. Environ. Sci. Pollut. Res. 25, trocoagulation. World Patent Application WO 2013/059964 A1.
32844–32850. doi:10.1007/s11356- 017- 8390- 1. Xin, S., Nin, C., Gong, Y., Ma, J., Bi, X., Jiang, B., 2018. A full-wave rectified alter-
Thiam, A., Zhou, M., Brillas, E., Sirés, I., 2014. Two-step mineralization of Tartrazine nating current wireless electrocoagulation strategy for the oxidative remedia-
solutions: Study of parameters and by-products during the coupling of electro- tion of As(III) in simulated anoxic groundwater. Chem. Eng. J. 351, 1047–1055.
coagulation with electrochemical advanced oxidation processes. Appl. Catal. B doi:10.1016/j.cej.2018.06.178.
Environ. 150–151, 116–125. doi:10.1016/j.apcatb.2013.12.011. Xu, H.Y., Yang, Z.H., Zeng, G.M., Luo, Y.L., Huang, J., Wang, L.K., Song, P.P., Mo, X.,
Thostenson, J.O., Mourouvin, R., Hawkins, B.T., Ngaboyamahina, E., Sellgren, K.L., 2014. Investigation of pH evolution with Cr(VI) removal in electrocoagulation
Parker, C.B., Deshusses, M.A., Stoner, B.R., Glass, J.T., 2018. Improved blackwa- process: Proposing a real-time control strategy. Chem. Eng. J. 239, 132–140.
ter disinfection using potentiodynamic methods with oxidized boron-doped di- doi:10.1016/j.cej.2013.11.008.
amond electrodes. Water Res 140, 191–199. doi:10.1016/j.watres.2018.04.022. Xu, H.Y., Yang, Z.H., Luo, Y.L., Zeng, G.M., Huang, J., Wang, L.K., Song, P.P., Yang, X.,
Timmes, T.C., Kim, H.C., Dempsey, B.A., 2010. Electrocoagulation pretreatment of 2015. A novel approach to sustain Fe0-electrocoagulation for Cr(VI) removal by
seawater prior to ultrafiltration: Pilot-scale applications for military water pu- optimizing chloride ions. Sep. Purif. Technol. 156, 200–206. doi:10.1016/j.seppur.
rification systems. Desalination 250, 6–13. doi:10.1016/j.desal.2009.03.021. 2015.09.074.
Trompette, J.L., Vergnes, H., 2009. On the crucial influence of some supporting elec- Xu, L., Cao, G., Xu, X., Liu, S., Duan, Z., He, C., Wang, Y., Huang, Q., 2017. Simultaneous
trolytes during electrocoagulation in the presence of aluminum electrodes. J. removal of cadmium, zinc and manganese using electrocoagulation: Influence of
Hazard. Mater. 163, 1282–1288. doi:10.1016/j.jhazmat.2008.07.148. operating parameters and electrolyte nature. J. Environ. Manage. 204, 394–403.
van Genuchten, C.M., Bandaru, S.R.S., Surorova, E., Amrose, S.E., Gadgil, A.J., Peña, J., doi:10.1016/j.jenvman.2017.09.020.
2016. Formation of macroscopic surface layers on Fe(0) electrocoagulation elec- Xu, T., Zhou, Y., Lei, X., Hu, B., Chen, H., Yu, G., 2019. Study on highly efficient Cr(VI)
trodes during an extended field trial of arsenic treatment. Chemosphere 153, removal from wastewater by sinusoidal alternating current coagulation. J. Envi-
270–279. doi:10.1016/j.chemosphere.2016.03.027. ron. Manage. 249, 109322. doi:10.1016/j.jenvman.2019.109322.
van Genuchten, C.M., Behrends, T., Kraal, P., Stipp, S.L.S., Dideriksen, K., 2018. Con- Yang, Z.H., Xu, H.Y., Zeng, G.M., Luo, Y.L., Yang, X., Huang, J., Wang, L.K., Song, P.P.,
trols on the formation of Fe(II,III) (hydr)oxides by Fe(0) electrolysis. Electrochim. 2015. The behavior of dissolution/passivation and the transformation of passive
Acta 286, 324–338. doi:10.1016/j.electacta.2018.08.031. films during electrocoagulation: Influences of initial pH, Cr(VI) concentration,
van Genuchten, C.M., Dalby, K.N., Ceccato, M., Stipp, S.L.S., Dideriksen, K., 2017. and alternating pulsed current. Electrochim. Acta 153, 149–158. doi:10.1016/j.
Factors affecting the Faradaic efficiency of Fe(0) electrocoagulation. J. Environ. electacta.2014.11.183.
Chem. Eng. 5, 4958–4968. doi:10.1016/j.jece.2017.09.008. Yasri, N.G., Gunaskearan, S., 2017. Electrochemical Technologies for Environmental-
Vasudevan, S., Kannan, B.S., Lakshmi, J., Mohanraj, S., Sozhan, G., 2011. Effects of Remediation. In: Anjum, N., Gill, S., Tuteja, N. (Eds.), Enhancing Cleanup ofEn-
alternating and direct current in electrocoagulation process on the removal of vironmental Pollutants. Springer Nature, Cham, Switzerland, pp. 5–73. doi:10.
fluoride from water. J. Chem. Technol. Biotechnol. 86, 428–436. doi:10.1002/jctb. 1007/978- 3- 319- 55423- 5_2.
2534. Yasri, N.G., Hu, J., Kibria, G., Roberts, E.P.L., 2020. In: Chernyshova, I., Ponnu-
Vasudevan, S., Lakshmi, J., 2012. Effect of alternating and direct current in an elec- rangam, S., Liu, Q. (Eds.). American Chemical Society„ Washington DC, pp. 167–
trocoagulation process on the removal of cadmium from water. Water Sci. Tech- 203. doi:10.1021/bk- 2020- 1348.ch006.
nol. 65, 353–360. doi:10.2166/wst.2012.859. Yet, L., 2011. Five Membered Ring Systems. Chem. Eng. Sci. 57, 217–257. doi:10.1016/
Vepsäläinen, M., 2012. Electrocoagulation in the treatment of industrial waters and S0959- 6380(11)22008- 7.
wastewaters PhD thesis. VTT Technical Research Centre of Finland. Yildiz, Y.Ş., Koparal, A.S., Keskinler, B., 2008. Effect of initial pH and supporting elec-
Vepsalainen, M., Selin, J., Rantala, P., Pulliainen, M., Sarkka, H., Kuhmonen, K., Bhat- trolyte on the treatment of water containing high concentration of humic sub-
nagar, A., Sillanpaa, M., 2011. Precipitation of dissolved sulphide in pulp and stances by electrocoagulation. Chem. Eng. J. 138, 63–72. doi:10.1016/j.cej.2007.
paper mill wastewater by electrocoagulation. Environ. Technol. 32, 1393–1400. 05.029.
doi:10.1080/09593330.2010.536790. Zhang, B., Ma, X.L., 2019. A review—Pitting corrosion initiation investigated by TEM.
Vik, E.A., Carlson, D.A., Eikum, A.S., Gjessing, E.T., 1984. Electrocoagulation of potable J. Mater. Sci. Technol. 35, 1455–1464. doi:10.1016/j.jmst.2019.01.013.
water. Water Res 18, 1355–1360. doi:10.1016/0 043-1354(84)90 0 03-4. Zhou, R., Liu, F., Wei, N., Yang, C., Yang, J., Wu, Y., Li, Y., Xu, K., Chen, X., Zhang, C.,
Wang, C.T., Chou, W.L., Kuo, Y.M., 2009. Removal of COD from laundry wastewater 2020. Comparison of Cr(VI) removal by direct and pulse current electrocoag-
by electrocoagulation/electroflotation. J. Hazard. Mater. 164, 81–86. doi:10.1016/ ulation: Implications for energy consumption optimization, sludge reduction
j.jhazmat.2008.07.122. and floc magnetism. J. Water Process Eng. 37, 101387. doi:10.1016/j.jwpe.2020.
Wellner, D.B., Couperthwaite, S.J., Millar, G.J., 2018. Influence of operating parame- 101387.
ters during electrocoagulation of sodium chloride and sodium bicarbonate solu-
tions using aluminium electrodes. J. Water Process Eng. 22, 13–26. doi:10.1016/
j.jwpe.2017.12.014.

25

You might also like