You are on page 1of 65

Journal Pre-proof

The role of the current waveform in mitigating passivation and enhancing


electrocoagulation performance: A critical review

Javad Abdollahi, Mohammad Reza Alavi Moghaddam, Sajjad Habibzadeh

PII: S0045-6535(22)03705-5
DOI: https://doi.org/10.1016/j.chemosphere.2022.137212
Reference: CHEM 137212

To appear in: ECSN

Received Date: 18 August 2022


Revised Date: 7 November 2022
Accepted Date: 8 November 2022

Please cite this article as: Abdollahi, J., Alavi Moghaddam, M.R., Habibzadeh, S., The role of the current
waveform in mitigating passivation and enhancing electrocoagulation performance: A critical review,
Chemosphere (2022), doi: https://doi.org/10.1016/j.chemosphere.2022.137212.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2022 Published by Elsevier Ltd.


CRediT authorship contribution statement

Javad Abdollahi: Conceptualization, Investigation, Literature review, Writing- original draft

preparation, Writing- reviewing and editing, Visualization. Mohammad Reza Alavi

Moghaddam: Conceptualization, Writing- reviewing and editing, Supervision. Sajjad

Habibzadeh: Conceptualization, Writing- reviewing and editing, Supervision.

of
ro
-p
re
lP
na
ur
Jo
Jo
ur
na
lP
re
-p
ro
of
1 The role of the current waveform in mitigating passivation and
2 enhancing electrocoagulation performance: A critical review

3 Javad Abdollahi a, Mohammad Reza Alavi Moghaddam a,*, Sajjad Habibzadeh b

a
4 Department of Civil & Environmental Engineering, Amirkabir University of Technology
5 (Tehran Polytechnic), Iran.

b
6 Department of Chemical Engineering, Amirkabir University of Technology (Tehran
7 Polytechnic), Iran.

f
8 * Corresponding author: alavim@yahoo.com (M. R. Alavi Moghaddam)

oo
9 Tel: +98-21-64543008; Fax: +98-21-66414213

pr
e-
Pr
al
u rn
Jo

1
10 Abstract

11 Electrocoagulation (EC) can be an efficient alternative to existing water and wastewater

12 treatment methods due to its eco-friendly nature, low footprint, and facile operation. However,

13 the electrodes applied in the EC process suffer from passivation or fouling, an issue resulting

14 from the buildup of poorly conducting materials on the electrode surface. Indeed, such

15 passivation gives rise to various operational problems and restricts the practical implementation

16 of EC on a large scale. Therefore, it has been suggested that using pulsed direct current (PDC),

17 alternating pulse current (APC), and sinusoidal alternating current (AC) waveforms in EC as

f
oo
18 alternatives to conventional direct current (DC) can help mitigate passivation and alleviate its

19
pr
associated detrimental effects. This paper presents a critical review of the impact of the current
e-
20 waveform on the EC process towards the capabilities of the PDC, APC, and AC waveforms in
Pr

21 de-passivation and performance enhancement while comparing them to the conventional DC.

22 Additionally, current waveform parameters influencing the surface passivation of electrodes


al

23 and process efficiency are elaborately discussed. Meanwhile, the performance of the EC
rn

24 process is evaluated under different current waveforms based on pollutant removal efficiency,
u

25 energy consumption, electrode usage, sludge production, and operating cost. The proper
Jo

26 current waveforms for treating various water and wastewater matrices are also explained.

27 Finally, concluding remarks and outlooks for future research are provided.

28 Keywords: Electrocoagulation; Passivation; Current waveform; Pulse current; Polarity

29 reversal; Alternating current

2
30 1. Introduction

31 Electrocoagulation (EC) has been praised recently as a robust and eco-sustainable

32 technology in water and wastewater treatment, capable of eliminating a broad spectrum of

33 pollutants. EC incorporates the merits and capabilities of coagulation, flotation, adsorption, as

34 well as electrochemistry (Chen et al., 2018; Nidheesh et al., 2021). As a result, it offers many

35 advantages over the conventional treatment methods and is characterized by an eco-friendly

36 nature, ease of operation, versatility, a small footprint, and a minimum requirement of

37 chemicals (Sefatjoo et al., 2020; Shokri and Fard, 2022). However, the EC process, which is

f
oo
38 often operated with direct current (DC), suffers from high energy and electrode consumption,

39
pr
plus electrode passivation (Garcia-Segura et al., 2017; Al-Raad and Hanafiah, 2021; Yasri et
e-
40 al., 2022). This latter issue discourages the large-scale applicability and commercialization of
Pr

41 EC and can cause various malfunctions over long-term operations, such as decreased coagulant

42 production, excessive energy consumption, and diminished treatment efficiency (Fu et al.,
al

43 2021; Oliveira et al., 2021; Yu et al., 2021). Accordingly, electrode passivation should be
rn

44 controlled or eliminated to overcome these adverse aftereffects and harness the full potential
u

45 of the EC process (Ibrahim et al., 2020).


Jo

46 Many approaches have been proposed to deal with passivation, among which changing

47 the current waveform has gained the most attention. In recent years, there has been a growing

48 interest in operating EC with pulsed direct current (PDC-EC), alternating pulse current (APC-

49 EC), and sinusoidal alternating current (AC-EC). Although these current waveforms can

50 potentially prevent electrode passivation, their impact on the overall EC performance is still

51 unclear, and sometimes contradictory findings have been reported. Research in this field has

52 been intensive during the last two decades, and various research papers have been published.

53 However, although a few reviews addressed electrode passivation and discussed

54 passivation mitigation approaches (Ingelsson et al., 2020; Al-Raad and Hanafiah, 2021; Moradi
3
55 et al., 2021), to the authors' knowledge, no review article has been published specifically

56 regarding the role of PDC, APC, and AC waveforms in reducing electrode passivation and

57 improving EC performance. Bearing this in mind, this review seeks to fill this gap.

58 Accordingly, this study will summarize the main current waveforms employed in the EC

59 process after outlining the fundamentals of EC and explaining the passivation phenomenon.

60 This will be followed by comparing PDC-EC, APC-EC, and AC-EC with conventional DC-

61 EC and providing an in-depth discussion regarding the impact of the current waveform on

62 passivation and process performance from the viewpoints of pollutant removal efficiency,

f
oo
63 energy requirements, electrode consumption, sludge production, and operating costs.

64

pr
Meanwhile, the impact of current waveform parameters on PDC-EC and APC-EC will be

discussed. The proper current waveforms for treating various water and wastewater matrices
e-
65

66 will also be explained. Finally, several perspectives for future research will be highlighted.
Pr

67
al

68 2. Fundamentals of the electrocoagulation process


rn

69 EC is an electrochemical technology that resembles chemical coagulation, except that


u

70 coagulant species are generated in situ by applying electric current to the sacrificial electrodes
Jo

71 (Behbahani et al., 2011a; Taheri et al., 2013; Nazlabadi et al., 2019). EC involves multiple

72 mechanisms operating synergistically to eliminate pollutants. When the electrodes are exposed

73 to electric current, the anode is oxidized into metal ions (Eq. (1)), while concurrently, at the

74 cathode, water is reduced into H2 gas and OH− ions (Eq. (2)). Then, metal ions and OH− transfer

75 into the solution and undergo spontaneous hydrolysis reactions (Eq. (3)), forming monomeric

76 and polymeric species that eventually transform into insoluble metal oxides, hydroxides, or

77 oxyhydroxides (herein collectively referred to as metal (oxy)hydroxide flocs) (Karimifard and

78 Alavi Moghaddam, 2018; Tegladza et al., 2021). These compounds are responsible for the

79 adsorption or entrapment of pollutants due to their large surface area and high adsorption

4
80 capacity (Taheri et al., 2012). The agglomerated flocs and pollutants can settle as sludge or

81 attach to hydrogen bubbles and float via electro-flotation (Behbahani et al., 2011b; Karamati

82 Niaragh et al., 2017; Nazlabadi and Alavi Moghaddam, 2017).

n+
83 M(s) → M(aq) + ne− (1)


84 2H2 O(l) + 2e− → H2(g) ↑ +2OH(aq) (2)

n+ +
85 M(aq) + nH2 O(l) → M(OH)n(s) + nH(aq) (3)

86 The amount of coagulant generated from metal anode material dissociation can be

f
oo
87 estimated by Faraday's law (Eq. (4)) (Mollah et al., 2004). The faradaic efficiency, which is the

88 ratio of the amount of anode material actually dissolved to the theoretical value expected from

89 pr
Faraday's law, is generally used as an appraisal index to evaluate the EC process performance
e-
90 (Müller et al., 2019).
Pr

𝐼𝑡𝑀𝑤
91 𝑚= (4)
al

𝑛𝐹
rn

92 where m represents the mass of anode dissolved (g), Mw is the specific molecular weight of
u

93 anode material (g mol−1), I is the applied current intensity (A), t is the electrolysis time (s), F
Jo

94 is the Faraday's constant (96485 C mol−1), and n is the number of electrons involved in the

95 anodic reaction.

96 The overall performance of EC depends on a multitude of factors. The main factors that

97 influence EC and the vectors through which EC performance could be affected are illustrated

98 in Fig. 1. Given the focus of this review on the current waveform and the passivation

99 phenomenon, Fig. 1 only depicts how these two factors interact with others, which will be

100 discussed in the following sections.

5
of
ro
-p
re
lP
na
ur
Jo

101

102 Fig. 1. Factors affecting EC performance.

6
103 3. Electrode passivation in the electrocoagulation process

104 Maintaining a steady and high faradaic efficiency during prolonged operation is essential

105 for reliable EC performance (Müller et al., 2019). However, the accumulation of solid materials

106 on the electrode in the form of a surface layer/film, an issue that is referred to by different terms

107 such as electrode passivation or fouling, can impede the mass transfer of metal ions into the

108 bulk solution due to their entrapment in the structure of the surface layer or lead to insufficient

109 metal ions production due to the occurrence of side reactions instead of electrode oxidation

110 (van Genuchten et al., 2016; Müller et al., 2019). Several factors can affect electrode

f
oo
111 passivation, such as solution composition, solution pH, current density, current waveform,

112
pr
reactor design, pollutant concentration, dissolved oxygen, and electrode material (Ingelsson et
e-
113 al., 2020; Al-Raad and Hanafiah, 2021; Chow and Pham, 2021; Guo et al., 2022).
Pr

114 Among these factors, solution composition is of great importance (Syam Babu et al.,

115 2021). The passivation rate in various water and wastewater matrices may vary dramatically
al

116 owing to the different types and amounts of pollutant species in their composition. Lower
rn

117 passivation rates are expected in tap water due to the low concentration of interfering ions
u
Jo

118 (Khandegar and Saroha, 2013), while in groundwater, the presence of Ca and Mg cations and

119 oxyanions, such as nitrate, carbonate, phosphate, and sulfate, can contribute to higher

120 passivation rates (van Genuchten et al., 2017; Chow et al., 2021; AlJaberi et al., 2022).

121 Additionally, natural pollutants in water, such as natural organic matter (NOM), have been

122 reported to intensify passivation (Mohora et al., 2014). However, the passivation rate could be

123 much higher and particularly problematic in various wastewater matrices because of the higher

124 contaminants and COD load (Sadeghi et al., 2022; Yasasve et al., 2022). Due to the diversity

125 of ionic species and contaminants causing passivation (van Genuchten et al., 2017), the

126 passivation rate can differ from one industrial wastewater to another. The passivation

7
127 mechanisms, detrimental effects of passivation on EC, and de-passivation approaches are

128 summarized in Fig. SM-1 and discussed in the following paragraphs.

129

130 3.1. Passivation mechanisms

131 Electrode passivation results from electrochemical and physicochemical mechanisms

132 that involve coagulants and aqueous-phase species reacting and precipitating on the electrode

133 surface (Yasri et al., 2022). Water electrolysis forms acidic and alkaline pH environments

134 around the anode and cathode, respectively. Therefore, ionic species in the electrolyte may

f
oo
135 deposit on the electrode surface if they exceed their solubility limitations (Ingelsson et al.,

136
pr
2020; Yasri et al., 2022). The following are the main reactions and mechanisms that contribute
e-
137 to the formation of the surface layer and increase passivation: (1) the precipitation of a portion

of metal (oxy)hydroxide flocs on the electrode surface (Timmes et al., 2010; Fu et al., 2021;
Pr

138

139 Yu et al., 2021; W. Zhang et al., 2021), (2) the bonding of oxyanions to metal ions and thus the
al

140 formation of the oxyanion-rich passivating film on the anode surface (Mansouri et al., 2011;
rn

141 Dubrawski et al., 2015; van Genuchten et al., 2017; Xu et al., 2017), (3) the precipitation of Ca
u

142 and Mg minerals on the cathode surface (Haldar and Gupta, 2020; Dutta et al., 2021; AlJaberi
Jo

143 et al., 2022; Dutta and Gupta, 2022), (4) the accumulation of contaminant residues on the

144 electrode surface (Eyvaz et al., 2009; Xu et al., 2015; Yang et al., 2015), and (5) the adsorption

145 of other particles in solution to the anode or cathode surface (Ingelsson et al., 2020). A

146 schematic of the mechanisms occurring during EC is illustrated in Fig. 2.

8
f
oo
pr
e-
Pr
al
u rn
Jo

147

148 Fig. 2. The schematic diagram of EC and the mechanisms involved in the process

149

150 3.2. Detrimental effects of passivation

151 Once the surface layer is formed on the electrode surface, this layer is expected to thicken

152 over time due to the decreased probability of metal ion migration through the layer to the bulk

153 solution (Cesar Lopes Geraldino et al., 2016). If the surface layer consists of poorly conducting

154 materials, the solid buildup can result in various operational concerns (Bandaru et al., 2020).

155 The passivated electrode is prone to pitting corrosion, resulting in structural failure and periodic

9
156 replacement at shorter intervals (Punckt et al., 2004; Dura and Breslin, 2019; Zhang and Ma,

157 2019; Trompette and Lahitte, 2021). Besides, electrode passivation makes the electrode surface

158 quasi-inert and impedes normal electron flow. This phenomenon subsequently increases ohmic

159 resistance and overpotential, resulting in excessive energy consumption and higher operating

160 costs (Madhavan and Antony, 2021; Payami Shabestar et al., 2021; Zhu et al., 2022). In

161 addition, at sufficiently high anodic overpotentials, the formation of passive film favors water

162 oxidation (Eq. (5)) rather than electrode oxidation, which reduces the amount of formed flocs

163 and diminishes pollutant removal efficiency markedly (Yang et al., 2015; Ashraf et al., 2019;

f
oo
164 Alkhatib et al., 2020). Moreover, severe passivation can cause the system to shut down (Yasri

165 et al., 2022).


pr
e-
+
166 2H2 O(l) → 4H(aq) + O2(g) ↑ +4e− (5)
Pr

167

168 3.3. Passivation mitigation or prevention approaches


al
rn

169 Since the formation and growth of the passivating layer on the electrode surface can

170 negatively affect the EC process, electrode de-passivation is needed to maintain the EC
u
Jo

171 performance over extended periods. The main strategies employed in EC to de-passivate

172 electrodes, along with their merits and drawbacks, are mentioned below:

173 (1) Mechanical or chemical cleaning of electrodes. Merits: effective de-passivation; no

174 need for additional equipment. Drawbacks: high labor expenses; requires system

175 shutdown (a challenge for automating EC); needs acid waste management (for

176 chemical cleaning) (Amrose et al., 2014; Bandaru et al., 2020).

177 (2) Adding aggressive ions such as Cl− to the electrolyte. Merits: effective in-situ de-

178 passivation; simple; no need for extra equipment. Drawbacks: risk of hazardous

179 byproducts; needs further treatment (Arroyo et al., 2009; Ingelsson et al., 2020).

10
180 (3) Increase the turbulent state by modifying the electrode shape, reactor geometry, or

181 flow velocity. Merits: effective in-situ de-passivation; no need for chemicals.

182 Drawbacks: not applicable for a batch system; requires a continuous pump (Timmes

183 et al., 2010; Ibrahim et al., 2020; Al-Raad and Hanafiah, 2021).

184 (4) Applying a super-gravity field. Merits: effective in-situ de-passivation; no need for

185 chemicals. Drawbacks: requires extra infrastructure; complex setup (Yu et al., 2021).

186 (5) Combining the EC process with the ultrasonic (Sono-EC). Merits: effective in-situ

187 de-passivation; no need for chemicals. Drawbacks: requires special infrastructure;

f
oo
188 expensive; may cause flocs to break down (He et al., 2016; Moradi et al., 2021).

189

pr
(6) Using current waveforms other than DC. Merits: in-situ de-passivation; complete

automatability; no complex design modification; no additional treatment phases; no


e-
190

191 need for chemicals. Drawbacks: requires additional electrical equipment (Chow and
Pr

192 Pham, 2021; Arabameri et al., 2022; Dong et al., 2022).


al

193 Although strategies 1 to 5 could be effective in passivation mitigation, they require either
rn

194 chemical addition, system termination, or special infrastructure, all of which considerably raise
u
Jo

195 the cost and complexity of the process. However, EC operation with current waveforms other

196 than DC (strategy 6) has become increasingly appealing in recent years due to the various

197 advantages outlined above. In addition, uniform electrode corrosion, reduced energy and

198 electrode consumption, lower sludge production, and cost-effectiveness make this approach

199 more desirable. Accordingly, this review focuses on this strategy and evaluates its capabilities

200 for electrode de-passivation and performance enhancement of EC.

201

202 4. Current waveforms in the electrocoagulation process

203 According to Fig. 3, the EC process can be classified into four main techniques based on

204 the current waveform adopted: direct current EC (DC-EC), pulsed direct current EC (PDC-

11
205 EC), alternating pulse current EC (APC-EC), and sinusoidal alternating current EC (AC-EC).

206 The electric current in DC-EC and PDC-EC is unidirectional, with the latter being

207 discontinuous. In contrast, the electric current in APC-EC and AC-EC is bidirectional; thus,

208 the current direction and polarity of electrodes alter intermittently. Most EC studies have used

209 DC, which has several drawbacks, such as high energy and electrode consumption, insufficient

210 residence time for flocculation, electrode inefficiency, and electrode passivation (Rajaei et al.,

211 2021). The PDC-EC, APC-EC, and AC-EC techniques have been proposed to overcome the

212 disadvantages of DC-EC. These techniques have drawn significant attention since they can

f
oo
213 improve performance by manipulating coagulant mass transfer, system resistance, coagulant

214

pr
species produced, and bubble gas generation (Karamati-Niaragh et al., 2019; Zhou et al., 2020;

Madhavan and Antony, 2021; Payami Shabestar et al., 2021; Arabameri et al., 2022).
e-
215
Pr

216 Several studies have operated the EC process with PDC, APC, and AC waveforms in the

217 last two decades. This review focuses on those studies that compared these waveforms to
al

218 conventional DC by considering different criteria, namely pollutant removal efficiency, energy
rn

219 usage, electrode consumption, sludge production, and operation cost, as well as research that
u
Jo

220 optimized current waveform parameters to fully exploit EC potential.

12
of
ro
-p
re
lP
na
ur
Jo

221

222 Fig. 3. EC techniques based on the current waveform: (a) DC-EC, (b) PDC-EC, (c) AC-EC, and (d) APC-EC

13
223 5. Pulsed direct current electrocoagulation (PDC-EC)

224 Periodic interruptions of direct current in EC have been demonstrated to help manage

225 electrode passivation and improve performance. In this technique, which is referred to by

226 different terms such as pulsed direct current EC (PDC-EC), pulse EC (PEC), or positive single

227 pulse current EC (PSPC-EC), the DC power supply is kept on for a given time and then

228 switched off for a specified time in a pulse period (Fig. 3(b)). Therefore, unlike in DC-EC, the

229 electrode reactions are discontinuous in PDC-EC (Ren et al., 2011). The duration of current

230 on- and off-times in this technique is compared to DC-EC. The duration of current on- and off-

f
oo
231 times in this technique can be adjusted by modifying the current waveform parameters, which

232
pr
would affect process performance. In the following, the role of current waveform parameters
e-
233 on PDC-EC is discussed, and the capability of this technique in reducing electrode passivation
Pr

234 and enhancing treatment performance is compared to DC-EC.

235
al

236 5.1. The impact of current waveform parameters on PDC-EC


rn

237 The waveform of every PDC is unique and can be characterized by (1) pulse frequency
u
Jo

238 and (2) duty cycle. Table SM-1 presents a summary of studies that explored the impact of these

239 parameters on PDC-EC.

240

241 5.1.1. Pulse frequency

242 The pulse frequency is the reciprocal of the single operating period of the pulse (f= 1/T)

243 and represents the number of pulses per unit of time (Zhou et al., 2020; Dong et al., 2022).

244 According to Table SM-1, the optimal pulse frequencies recommended in the literature range

245 from 2.78 mHz (6 min pulse period) for treating synthetic dye wastewater with Al/Fe electrode

246 pairs (Akhbarati et al., 2017) to 10 kHz (100 μs pulse period) for COD removal from municipal

247 wastewater with Al electrodes (Chen et al., 2011). The broad scope of suggested pulse

14
248 frequencies could be attributed to differences in criteria considered in PDC-EC optimization

249 and operating conditions, such as solution chemistry, pollutant type, and electrode material.

250 Generally, a low-frequency PDC can passivate the electrode surface (Yang et al., 2020),

251 while a high pulse frequency can impede concentration polarization and passive layer

252 formation within a short pulse period, resulting in enhanced treatment efficiency (Zhou et al.,

253 2020). The high-frequency PDC has also been claimed to have a stronger penetrability to

254 puncture passivating film to help produce more coagulant agents (Zhou et al., 2020; An et al.,

2022). However, most studies reported that the pulse frequency had a non-linear impact on

f
255

oo
256 treatment efficiency, and increasing this variable did not necessarily enhance pollutant removal

257
pr
efficiency (Chen et al., 2011; Ren et al., 2011; Zheng, 2017; Yang et al., 2020; Zhou et al.,
e-
258 2020; An et al., 2022; Dong et al., 2022). This behavior could be explained by the concept that
Pr

259 the electrical double-layer in EC functions as a capacitor (Hasani et al., 2019), so at very high

260 pulse frequencies, the pulse width and current off-time can become shorter than the charge-
al

261 discharge time of the electrode double-layer, causing PDC-EC to fail to eliminate passivation
rn

262 and leading to lower pollutant removal efficiency (Ren et al., 2011). Thus, the current on-time
u
Jo

263 in a pulse period should be long enough for the electrical double-layer to be fully charged, and

264 the off-time should be so long that the double-layer has enough time to discharge (Hasani et

265 al., 2019). As a result, depending on the experimental conditions and the scope of investigated

266 frequencies, changes in pulse frequency can result in different scenarios in terms of passivation

267 mitigation and pollutant removal efficiency.

268

269 5.1.2. Duty cycle

270 The duty cycle (γ), which is the ratio of pulse width (ton) to the pulse period (T= ton +

271 toff), distinguishes between the PDC and conventional DC waveforms (Wang and Liu, 2017)

272 and directly affects energy consumption and faradaic efficiency in PDC-EC (Chen et al., 2011).

15
273 With a very low duty cycle, only a tiny proportion of the time in a pulse period is spent on

274 electrification, which reduces metal ion production and treatment efficiency (Dong et al.,

275 2022). However, a low duty cycle can decrease concentration polarization and electrode

276 passivation, resulting in higher pollutant removal efficiency (Yang et al., 2020). On the other

277 hand, increasing the duty cycle can increase metal ion generation and floc formation, improving

278 pollutant removal effectiveness. However, due to increased electrode dissolution at higher duty

279 cycles, the metal ions can be difficult to diffuse into the bulk solution, which results in

280 concentration polarization and electrode passivation, diminishing treatment efficiency (Yang

f
oo
281 et al., 2020; An et al., 2022; Zhang et al., 2022). Some studies also claimed that the duty cycle

282

pr
might affect floc structure and, thus, treatment efficacy (Dong et al., 2022).
e-
283 Furthermore, according to Eq. (6), the energy consumption in PDC-EC is dependent on
Pr

284 the duty cycle, and lower values of this variable may favor more energy conservation, while at

285 higher duty cycles, PDC tilts toward DC, leading to higher energy usage (Ren et al., 2011;
al

286 Wang and Liu, 2017). However, it should be noted that operation time may be prolonged with
rn

287 a low duty cycle to achieve similar removal efficiency to that of a high duty cycle, increasing
u
Jo

288 energy consumption (Zhou et al., 2020). Considering all the above, the majority of authors

289 reported that duty cycles in the range of 0.3–0.6 were favorable for optimum PDC-EC

290 performance (see Table SM-1).

291 𝐸 = 𝑈𝐼𝑡𝛾 2 (6)

292 where E is the energy consumption, U is the peak voltage, I is the peak current, t is the reaction

293 time, and γ is the duty cycle.

16
294 5.2. Comparison of PDC-EC and DC-EC techniques

295 Table 1 compiles the studies that compared PDC-EC and DC-EC techniques for various

296 pollutants and different electrode materials (e.g., aluminum, iron, and stainless steel) in terms

297 of electrode passivation and treatment performance. In most studies, the PDC-EC technique

298 has proven to be more efficient in abating pollutants than DC-EC (Ren et al., 2011; Akhbarati

299 et al., 2017; Wang and Liu, 2017; Zheng, 2017; H. Nguyen, 2020; Nguyen et al., 2020; Que et

300 al., 2021; Zhang et al., 2022). Higher treatment efficiency in PDC-EC has been attributed to

301 lower concentration polarization and diminished electrode passivation, as metal ions can

f
oo
302 diffuse into the solution and their concentration near the electrodes returns to the initial value

303
pr
when the current is turned off (Ren et al., 2011; Dong et al., 2022). It has also been reported
e-
304 that the PDC waveform can provoke mass transport by transient disruption of

attracting/repelling forces caused by electrodes, allowing ions or molecules such as metal


Pr

305

306 cations and OH− to readily contact each other (Chen et al., 2011). Besides, each pulse period's
al

307 current break (current off-time) can contribute to higher efficiency by giving the coagulants
rn

308 sufficient time to complete coagulation/flocculation reactions (Oliveira et al., 2021). For
u

309 instance, it has been revealed that employing the PDC waveform in the electrochemical process
Jo

310 using the Fe anode to eliminate Cr(VI) facilitated the Fe ion formation and led to more Fe ion

311 generation than the DC system (Zhou et al., 2019). As such, the concentration of Fe ions was

312 significantly enhanced by 707%, and the Cr concentration was reduced by 96% in the system

313 driven with the PDC compared to the system with the DC, demonstrating the more exhaustive

314 reactions when the PDC waveform was applied (Zhou et al., 2019).

17
315 Table 1. The summary of the reported studies on the comparison of PDC-EC with DC-EC

Process performance
Opt. PDC
Pollutant and
Anode/cathode waveform I, U, or j Removal Sludge Major outcomes References
solution Energy consumption Electrode consumption Operating
parameters efficiency production
(ENC) (ELC) cost (OC)
(RE) (SLP)
Al/Al COD in real f= 10 kHz PDC-EC: PDC-EC: PDC-EC: – – – Prevented ELP and uniform electrode Que et al.
municipal WW γ= 0.01 50 V 80% COD 2.8 kWh m−3 (for 68% corrosion in PDC-EC; 12% higher RE (2021)
DC-EC: DC-EC: COD) and 12.5% reduced ENC in PDC-EC
5V 68% COD DC-EC: than in DC-EC
3.2 kWh m−3
Al/Al COD in real f= 10 kHz PDC-EC: PDC-EC: PDC-EC: – – – PDC-EC: 17% higher RE and 24% less Nguyen et
textile WW γ= 0.01 45 V 77% COD 1.6 kWh m−3 (for 60% ENC than DC-EC al. (2020)

of
DC-EC: DC-EC: COD)
4V 60% COD DC-EC:

ro
2.1 kWh m−3

Al/Al, Fe/Fe* BH and COD in f= 1 kHz 194 A m−2 PDC-EC: PDC-EC: PDC-EC: – – Avoided ELP and more uniform Ren et al.

-p
synthetic and real γ= 0.3 90.1% BH 0.73 Wh (kg COD)−1 5.1 kg Fe (kg COD)−1 corrosion in PDC-EC; Higher removal (2011)

re
pharmaceutical 62.6% COD DC-EC: DC-EC: rates, 91% less ENC, and 11% lower
factory WW 8.20 Wh (kg COD)−1 5.7 kg Fe (kg COD)−1 ELC in PDC-EC than in DC-EC; Higher

lP
REs with Fe than with Al in PDC-EC
Al/Al, Fe/Fe* Dibutyl phthalate γ= 0.6 150 A m−2 PDC-EC: PDC-EC: PDC-EC: – – Greater passivation and lower RE with Wang and
in synthetic and 75% DBP 860 J (ppm DBP)−1 0.35 mg (ppm DBP)−1 Al than with Fe in PDC-EC; 41% lower Liu (2017)

na
real plastic DC-EC: DC-EC: ELC, 39% reduced ENC, and higher RE
factory WW 1420 J (ppm DBP)−1 0.59 mg (ppm DBP)−1 in PDC-EC than in DC-EC

Al/Al, Fe/Fe, Antimony in f= 0.4 kHz 15 V PDC-EC:


ur
PDC-EC: – – – PDC-EC: lower ENC than DC-EC for Dong et al.
Jo
Al/Fe, Fe/Al* synthetic WW γ= 0.5 97.4 antimony 54.8% less ENC than almost the same RE (2022)
DC-EC: DC-EC
97.2 antimony

Al/Fe Acid Blue 113 in f= 2.78 mHz 0.68 A PDC-EC: PDC-EC: – – – PDC-EC: 86% less ENC and a slightly Akhbarati et
synthetic WW γ= 0.5 99.66% dye 1.2 kWh m−3 higher RE than DC-EC al. (2017)
DC-EC: DC-EC:
98% dye 8.5 kWh m−3

Fe/Fe Direct Scarlet f= 0.2 kHz 168 A m−2 PDC-EC: PDC-EC: PDC-EC: – – PDC-EC: 51% lower ENC, 5% less ELC, Chen et al.
4BS and COD in γ= 0.1 99.61% color 1.110 J (mg COD)−1 0.187 kg Fe (kg and slightly lower REs than DC-EC (2011)
synthetic WW 91.46% COD DC-EC: COD)−1
DC-EC: 2.273 J (mg COD)−1 DC-EC:
99.72% color 0.197 kg Fe (kg
92.75 COD COD)−1

18
Process performance
Opt. PDC
Pollutant and
Anode/cathode waveform I, U, or j Removal Sludge Major outcomes References
solution Energy consumption Electrode consumption Operating
parameters efficiency production
(ENC) (ELC) cost (OC)
(RE) (SLP)
Fe/Fe COD, oil, f= 3 kHz PDC-EC: PDC-EC: PDC-EC: PDC-EC: PDC-EC: – Lessened ELP in PDC-EC; 76% lower Zheng
turbidity, TSS, γ= 0.3 350 A m−2 98.3% COD 0.19 kWh (kg COD)−1 3.1 kg Fe (kg COD)−1 5.3 kg (kg ENC, 9% less ELC, 4% reduced SLP, (2017)
and PAM in real DC-EC: 99.0% oil DC-EC: DC-EC: COD)−1 and slightly higher REs in PDC-EC than
oilfield produced 100 A m−2 DC-EC: 0.78 kWh (kg COD)−1 3.4 kg Fe (kg COD)−1 DC-EC: in DC-EC
water 94.7% COD 5.5 kg (kg
96.2% oil COD)−1
Fe/Fe Cadmium in f= 2 kHz PDC-EC: PDC-EC: PDC-EC: – – – Reduced ELP and larger surface area of Yang et al.
synthetic WW γ= 0.4 16 A m−2 99.92% Cd 0.400 kWh m−3 flocs in PDC-EC than in DC-EC; 44% (2020)
DC-EC: DC-EC: DC-EC: less ENC in PDC-EC than in DC-EC for
40 A m−2 99.74% Cd 0.714 kWh m−3 similar RE

of
Fe/Fe COD in real f= 10 kHz PDC-EC: PDC-EC: PDC-EC: – – – Uniform Fe corrosion in PDC-EC; H. Nguyen
γ= 0.01 1.7 kWh m−3 (for 72%

ro
municipal WW 50 V 82% COD Enhanced adsorption capacity of PDC- (2020)
DC-EC: DC-EC: COD) EC's flocs than those of DC-EC; 10%

-p
5V 72% COD DC-EC: higher RE and 39% less ENC in PDC-
2.8 kWh m−3 EC than in DC-EC

re
Fe/Fe Hardness and f= 2 kHz 120 A m−2 PDC-EC: PDC-EC: – – PDC-EC: Slightly less TH removal and 8% lower An et al.
turbidity in real γ= 0.6 23.35% TH 0.81 kWh m−3 0.15 $ m−3 ENC in PDC-EC than in DC-EC (2022)

lP
coal gasification 28.57% turbidity DC-EC:
gray water DC-EC: 0.88 kWh m−3
24.46% TH

na
22.86% turbidity
Fe/Fe Tetracycline in γ= 0.6 0.2 A PDC-EC: PDC-EC: PDC-EC: – PDC-EC: Improved adsorption efficiency in PDC- Zhang et al.

ur
synthetic and real 96.69% TC 0.11 kWh m−3 0.2042 kg Fe m−3 0.011 $ m−3 EC compared to DC-EC; Almost similar (2022)
livestock WW DC-EC: RE in both techniques
Jo
96.16% TC
Al/Al*, SS/SS Trimethoprim f= 120 Hz 120 A m−2 PDC-EC: PDC-EC: PDC-EC: – PDC-EC: Passivation effect of AMX and TMP Oliveira et
and amoxicillin γ= 0.7 8.2% TMP 1.65 kWh m−3 0.21 kg Al m−3 0.6 $ m−3 toward Al/Al. Corrosion effect of AMX al. (2021)
in real municipal 21.9% AMX toward SS/SS. 30% lower ENC in PDC-
WW 9.38% TPM + EC than in DC-EC
AMX
Al/SS Pharmaceuticals f= 0.67 mHz 3 A m−2 PDC-EC: ~15% PDC-EC: PDC-EC: – PDC-EC: Limited ELP in PDC-EC; 96% less ENC, Ensano et
in real municipal γ= 0.2 less REs than 0.21 kWh m−3 31.88 g m−3 0.1 € m−3 and 15% lower RE in PDC-EC than in al. (2019)
WW DC-EC DC-EC
Al/SS Sulfate in real f= 1.11 mHz 65 A m−2 PDC-EC: – – – – PDC-EC: 26% lower RE than DC-EC Rodrigues et
mine-impacted γ= 0.33 45.29% sulfate al. (2020)
water DC-EC:
70.95% sulfate
* The best electrode combination
316
I: Current; U: Voltage; j: Current Density; WW: Wastewater; ELP: Electrode passivation; SS: Stainless steel
317
19
318 Furthermore, it was found that the pulse current significantly impacts pollutant removal

319 in EC by affecting the adsorption efficiency (Zhang et al., 2022). Periodic terminating and

320 restarting the current in PDC-EC has been shown to help generate flocs with better adsorption

321 characteristics than those produced in DC-EC (H. Nguyen, 2020; Nguyen et al., 2020; Yang et

322 al., 2020; Zhou et al., 2020). For example, it was found that the flocs produced by PDC-EC

323 had a larger surface area for cadmium (Cd) adsorption and formed larger particles compared

324 to DC-EC (Yang et al., 2020). Due to the longer electrification time in DC-EC, it was also

325 demonstrated that a large amount of oxygen was produced due to water oxidation, leading to

f
oo
326 lepidocrocite (γ–FeOOH) formation. On the contrary, the green rust–Cl was produced in PDC-

EC in the absence of dissolved oxygen and the presence of Cl−, with the latter having a superior
327

pr
Cd removal property to the former (Yang et al., 2020). In contrast to these findings, a few
e-
328

329 studies have reported poorer performance of PDC-EC in removing pollutants than DC-EC
Pr

330 (Chen et al., 2011; Ensano et al., 2019; Rodrigues et al., 2020). For instance, it was shown that
al

331 the removal efficiency of pharmaceuticals was around 15% lower in the PDC-EC (5 min on/20
rn

332 min off) compared to the DC-EC (Ensano et al., 2019). The lower treatment efficiency in PDC-
u

333 EC was explained by the insufficient liberation of Al3+ species, which are responsible for the
Jo

334 charge neutralization of suspended pollutants and the adsorption of dissolved organic matter

335 (Ensano et al., 2019).

336 Regarding energy consumption, all the reviewed studies indicated that the PDC-EC

337 technique was more efficient than DC-EC. This behavior can be attributed to the shorter

338 electrification time in PDC-EC (Ren et al., 2011; Oliveira et al., 2021). Besides, reduced

339 electrode passivation in PDC-EC can lead to lower cell voltage and, thus, decreased energy

340 consumption. Also, due to the discontinuous reactions of converting electrical energy to

341 chemical energy in PDC-EC, the energy conversion in this technique could be more efficient

342 than that of DC-EC, and all the electrical energy can be concentrated on electrode corrosion

20
343 rather than diverted to water electrolysis for oxygen generation. As a result, efficient energy

344 utilization in PDC-EC can contribute to energy conservation (Chen et al., 2011). An inspection

345 of the articles reported in Table 1 reveals that PDC-EC can save up to 96% energy compared

346 to DC-EC for the same treatment efficiency (Ensano et al., 2019).

347 It has also been demonstrated that PDC-EC consumes fewer electrodes (Chen et al.,

348 2011; Ren et al., 2011; Wang and Liu, 2017; Zheng, 2017) and generates less sludge than DC-

349 EC (Zheng, 2017; Zhou et al., 2020). In PDC-EC, the electrode corrosion stops during the

current breaks, leading to reduced electrode utilization and lower sludge production (Ren et al.,

f
350

oo
351 2011). According to the data presented in Table 1, changing the current from DC to PDC can

352
pr
cut off electrode consumption and sludge production by up to 41% (Wang and Liu, 2017) and
e-
353 30% (Zhou et al., 2020), respectively. Unfortunately, among the reviewed publications, no
Pr

354 study has compared the operating costs of the PDC-EC and DC-EC techniques; however, due

355 to the decreased electrode and energy consumption and reduced sludge production in PDC-EC,
al

356 the overall operational cost of this technique is expected to be less than that of DC-EC.
rn

357
u
Jo

358 6. Alternating pulse current electrocoagulation (APC-EC)

359 EC with alternating pulse current, referred to as APC-EC, has drawn the most attention

360 among the EC techniques to restrain passive layer growth and improve process performance.

361 This technique involves intermittently reversing the current direction, which is generally

362 carried out by using an APC power supply (Xu et al., 2018; J. Zhang et al., 2021) or plugging

363 a programmable time relay into a DC power source (Eyvaz et al., 2009; Yang et al., 2015; Yasri

364 et al., 2022), allowing the electrodes to serve either as the anode or cathode for the prescribed

365 time intervals. Most studies in APC-EC just reversed the electrode's polarity, which is often

366 referred to as polarity reversal EC, or PR-EC (Fig. 3(d1)) (Eyvaz et al., 2009; Secula et al.,

367 2013; Donneys-Victoria et al., 2020; Chow et al., 2021). However, some researchers

21
368 interrupted the electric current in between polarity reversals (Fig. 3(d2)) (Jiang et al., 2018;

369 Xin et al., 2018) or employed varying numbers of pulses during the original and reversed

370 polarity (Fig. 3(d3)) (Mao et al., 2008; Xu et al., 2017; J. Zhang et al., 2021). The duration and

371 sequence of current on/off times and polarity reversal can be adjusted by manipulating the

372 current waveform parameters, influencing APC-EC performance. In this context, several

373 studies have been conducted, the noteworthy findings of which are reviewed in the section

374 below:

375

f
oo
376 6.1. The impact of current waveform parameters on APC-EC

377
pr
The following current waveform parameters have been found to affect APC-EC: (1)
e-
378 polarity reversal time, (2) pulse frequency, and (3) duty cycle. The polarity reversal time is

unique to the APC waveform, whereas pulse frequency and duty cycle are the common
Pr

379

380 properties of pulse current waveforms (PDC and APC). Properly controlling the APC
al

381 waveform parameters can help mitigate electrode passivation and assure high process
rn

382 performance.
u

383
Jo

384 6.1.1. Polarity reversal time

385 The polarity reversal time (PR time, Trev), which is the operation time before the

386 electrode's polarity is inverted, is considered a crucial element in APC-EC; as a result, the

387 proper selection of the PR time is critical. Table SM-2 compiles the studies that explored the

388 impact of PR time on the APC-EC technique and highlights the optimal PR times and the

389 authors' findings. In most studies, the duration of the forward and reversed polarities was the

390 same and equal to half of the cycle period (Pi et al., 2014; Yang et al., 2015; Fekete et al., 2016;

391 Betancor-Abreu et al., 2019; Maher et al., 2019; Ashraf et al., 2021; Chow et al., 2021; Yasri

392 et al., 2022). Meanwhile, some researchers operated the process with asynchronous polarity

22
393 reversal for hybrid electrode pairs (e.g., Fe/Al and Al/Fe), where electrodes serve as the anode

394 for different lengths of time (Xu et al., 2017; Fu, 2018; Jiang et al., 2018).

395 According to Table SM-2, the suggested PR times for optimal APC-EC performance

396 were too diverse, and different studies proposed a broad scope of optimal PR times, with the

397 shortest being 4 s for treating oily bilge water (Bian et al., 2019) and the longest being 30 min

398 for arsenic and NOM removal (Mohora et al., 2014). From Table SM-2, it can be inferred that

399 the optimal PR time for high APC-EC performance is dependent on operating conditions. For

example, the ideal PR time has been shown to vary for different electrode materials. Chow and

f
400

oo
401 Pham (Chow and Pham, 2021) found that a 10 min PR time gave the greatest balance of

402
pr
electrode de-passivation, high faradaic efficiency, and low energy consumption when using
e-
403 aluminum electrodes, yet it lowered neither passivation nor energy usage in the case of iron
Pr

404 electrodes. Besides, different PR times were found to have a contradictory impact on the

405 removal efficiency of various pollutants. For instance, it was shown that a 2 min PR time was
al

406 best for eliminating organic matters, but an 8 min PR time was ideal for removing chloride ions
rn

407 (Donneys-Victoria et al., 2020). Moreover, the suggested PR times vary in systems with
u
Jo

408 different electrolyte compositions since, for example, oxyanions such as sulfate and carbonate

409 can contribute to passivation while Cl− ions can de-passivate electrodes by inducing pitting

410 corrosion (Timmes et al., 2010; Dubrawski et al., 2015; van Genuchten et al., 2017; Wellner et

411 al., 2018; W. Zhang et al., 2021). As a result, it is not straightforward to directly compare the

412 results of previous studies due to the varying nature of passivating films and differences in

413 operating conditions such as electrode material, pollutant type, and electrolyte composition.

414 However, the optimal PR time should prevent passivation and ensure adequate coagulant

415 generation since high faradaic efficiency is a prerequisite for high APC-EC performance

416 (Chow et al., 2021).

23
417 From a mechanistic point of view, after each polarity reversal, three main mechanisms

418 have been reported to occur: (1) inversion of local pH around electrodes; (2) occurrence of side

419 reactions; and (3) charging of the electrode double-layer (Chow et al., 2021; Yasri et al., 2022).

420 The inversion of local pH around electrodes could be detrimental or beneficial to process

421 performance. For example, upon the reaction of metal ions generated in the acidic pH at the

422 anode with OH− produced in the basic pH at the cathode (the same electrode that served as the

423 anode before the polarity change), metal (oxy)hydroxides can form (Chow and Pham, 2021).

424 In addition to serving as coagulants, these species could precipitate on both the anode and

f
oo
425 cathode and contribute to passivation (van Genuchten et al., 2017). Thus, before the polarity is

426

pr
reversed, adequate time should be given for these species to transfer into the solution rather

than accumulate on the electrodes. However, the formation of high acidic/alkaline pH near Al
e-
427

428 electrodes after each polarity reversal can promote non-faradaic chemical corrosion of both the
Pr

429 anode and cathode (Fekete et al., 2016; Nidheesh et al., 2022), which aids in detaching the
al

430 precipitates from the electrode surface, removing the passive layer (Chow and Pham, 2021).
rn

431 The electrochemical side reactions can divert the electric current away from the metal
u
Jo

432 corrosion reaction and negatively affect process performance. Side reactions will likely occur

433 after polarity reversal by reactive species at the electrode interface, especially when Fe

434 electrodes are used. The Fe2+ and Fe3+ species generated from the oxidation of the Fe electrode

435 could be reduced and electrodeposited as Fe(0) and, in the subsequent reversal, oxidized again

436 (van Genuchten et al., 2016). Accordingly, fewer Fe species can transfer to the solution over

437 consecutive cycles, which lowers process efficiency. In contrast, Al3+ species have a relatively

438 low reduction potential compared to water hydrolysis, which prevents them from

439 electrodepositing as Al(0); thus, more Al ions can transfer into the solution (Ingelsson et al.,

440 2020). Other possible side reactions that might diminish process efficiency are Cl− and H2O

441 oxidation and the oxidation of H2 gas produced in the prior reversal cycle when the electrode

24
442 served as a cathode (Dubrawski et al., 2015; Ashraf et al., 2019). Charging of the electrode

443 double-layer is another mechanism that could impact treatment efficiency. The time required

444 for charging the double-layer is estimated to be about 2–100 ms for common electrolyte

445 conductivities of 1–20 mS cm−2 in the EC process, where higher conductivity shortens charging

446 time (Ingelsson et al., 2020). Since the main electrochemical reactions may not occur during

447 this time, shorter PR times than the charging time of the double-layer should be avoided.

448 Given all three mechanisms discussed above, under optimal operating conditions, the

application of the APC waveform could effectively mitigate passivation and improve EC

f
449

oo
450 efficacy, especially when Al electrodes were applied. However, the Fe electrode exhibits more

451
pr
complex electrochemical behavior than the Al electrode, which may explain the conflicting
e-
452 findings when Fe electrodes were employed in APC-EC. According to Fig. 4, longer PR times
Pr

453 were typically applied when the Fe electrode was employed compared to the Al electrodes. The

454 suggested PR times in various studies have been mostly longer than 4 min for Fe electrodes,
al

455 while PR times in the 1–5 min range have been recommended for Al electrodes.
u rn
Jo

25
f
oo
pr
e-
Pr
al
u rn
Jo

456

457 Fig. 4. The range of PR times investigated in various research and the optimal PR times

458 recommended for Al and Fe electrodes

26
459 6.1.2. Pulse frequency

460 The pulse frequency (f) is another essential property of the APC waveform that influences

461 electrode passivation and coagulant generation, thus affecting pollutant removal effectiveness

462 (Xu et al., 2021). The pulse frequency can also be adjusted to control APC-EC energy usage,

463 electrode consumption, and sludge generation. A few studies have been conducted to determine

464 the impact of pulse frequency on APC-EC performance, and conflicting results have been

465 reported in the literature (Xu et al., 2017, 2021; J. Zhang et al., 2021). The reviewed studies

466 showed that changes in pulse frequency result in varying effects on performance, so the

f
oo
467 recommended frequencies for optimal APC-EC performance varied greatly from 1 Hz for

468

pr
removing Cu2+ and COD in Fe APC-EC (Xu et al., 2021) to 5 kHz for eliminating Zn2+ and

Mn2+ from smelting wastewater using Fe/Al electrode pairs (Xu et al., 2017). These findings
e-
469

470 indicate that the optimal pulse frequency range for favorable APC-EC performance is still
Pr

471 unclear, and thus, further research is required to find the relationship between the tested
al

472 frequencies, operational conditions, and process performance.


rn

473
u

474 6.1.3. Duty cycle


Jo

475 The duty cycle plays a pivotal role in APC-EC and, similar to PDC-EC, influences the

476 energy usage and the number of dissolved metal cations from the electrodes. A handful of

477 studies investigated the impact of the duty cycle on APC-EC and optimized this variable to

478 achieve high treatment performance (Asaithambi et al., 2016, 2021; Xu et al., 2017; J. Zhang

479 et al., 2021). The reviewed studies reported that duty cycles of around 0.3–0.5 allowed the best

480 balance of electrode de-passivation, enhanced pollutant removal efficiency, and minimized

481 energy usage. In most cases, the duty cycle had a non-linear impact on pollutant removal

482 efficiency, such that by increasing the duty cycle, the pollutant removal efficiency first

483 increased and then decreased (Asaithambi et al., 2016, 2021; J. Zhang et al., 2021).

27
484

485 6.2. Comparison of APC-EC and DC-EC techniques

486 Table 2 summarizes the studies that compared the APC-EC technique to the conventional

487 DC-EC for treating various contaminants under different operating conditions. According to

488 Table 2, in most studies, APC-EC outperformed DC-EC in removing pollutants and

489 maintaining high treatment effectiveness over consecutive cycles (Mao et al., 2008; Eyvaz et

490 al., 2009; Pi et al., 2014; Ozyonar and Karagozoglu, 2015; Eyvaz, 2016; Jiang et al., 2018; Xin

491 et al., 2018; Alimohammadi et al., 2019; Donneys-Victoria et al., 2019; Asaithambi et al., 2021;

f
oo
492 Xu et al., 2021; J. Zhang et al., 2021; Yasri et al., 2022). Although improved efficiency in APC-

493
pr
EC is often claimed to be the direct consequence of reduced passivation, only a few authors
e-
494 (Mao et al., 2008; Chow et al., 2021; Chow and Pham, 2021; Yasri et al., 2022) analyzed the
Pr

495 properties of the electrode surface through material characterization methods.

496 The electrochemical reactions sweep between anode and cathode due to polarity reversal
al

497 in APC-EC, so gas formation and associated bubbling as well as electrode corrosion inverse
rn

498 intermittently, leading to a well-mixed situation in the EC cell, which can decrease the buildup
u
Jo

499 of the electrode passivating layer (Madhavan and Antony, 2021). Following the polarity

500 reversal, H2 gas generation at the cathode (the electrode that functioned as the anode before the

501 polarity change) can detach passivating layers formed on the electrode surface (Bian et al.,

502 2019; Chow et al., 2021). Meanwhile, the passive layers on the anode (the electrode that

503 formerly acted as the cathode) can be removed by the corrosion of the underneath electrode

504 (Bian et al., 2019; Chow et al., 2021). Besides, when the alkaline environment at the cathode

505 turns acidic after polarity reversal, the materials precipitated on the electrode surface may

506 dissolve, decreasing the formation rate of passive layers (Fekete et al., 2016; Chow et al., 2021;

507 Syam Babu et al., 2021; Yasri et al., 2022). Reduced passivation improves the EC's capability

508 to generate coagulants and enhances pollutant removal efficiency.

28
509 Table 2. The summary of the reported studies on the comparison of APC-EC with the DC-EC technique

Process performance
Opt. APC
Pollutant and
Anode/cathode waveform I, U, or j Removal Electrode Major outcomes References
solution Energy consumption Sludge production Operating
parameters efficiency consumption
(ENC) (SLP) cost (OC)
(RE) (ELC)
Al/Al COD in – 2A APC-EC: higher APC-EC: 30% less – – – Prevented ELP and uniform Al Mao et al.
synthetic WW RE removal than ENC than DC-EC corrosion in APC-EC; Higher RE (2008)
DC-EC and lower ENC in APC-EC than in
DC-EC
Al/Al Disperse and Trev= 150 s 105 A m−2 APC-EC: – – – APC-EC: Reduced ELP in APC-EC; Reduced Eyvaz et al.
reactive dyes 57–89% TOC 1.3–2.61 $ REs over time in DC-EC; Higher (2009)
in synthetic 63–99.9% dye (kg TOC)−1 REs in APC-EC than in DC-EC;
WW DC-EC: DC-EC: Lower OC in APC-EC than in DC-

of
49–81% TOC 1.43–3.43 $ EC
59–96% dye (kg TOC)−1

ro
Al/Al Silica in tap Trev= 900 s 48 A m−2 APC-EC: – APC-EC: – – ELP occurred in both techniques; Gelover-

-p
water 64.3% silica 0.0373 kg m−3 Slightly reduced RE and 6% higher Santiago et al.
DC-EC DC-EC: ELC in APC-EC than in DC-EC (2012)

re
66.5% silica 0.0353 kg m−3

1850 A m−2 –

lP
Al/Al Methyl Orange Trev= 15 s APC-EC: APC-EC: APC-EC: APC-EC: Delayed ELP in APC-EC; 15% Pi et al. (2014)
in synthetic 97% color 44 kWh (kg MO)−1 4.1 kg (kg MO)−1 17.2 kg (kg MO)−1 increased RE, 20% reduced ENC,
WW DC-EC: DC-EC: DC-EC: DC-EC: 11% higher ELC, and 6% less SLP

na
82% color 55 kWh (kg MO)−1 3.7 kg (kg MO)−1 18.3 kg (kg MO)−1 in APC-EC than in DC-EC

Al/Al TSS, turbidity, Trev= 300 s 1.2 V APC-EC & DC- APC-EC/biochar: – – – Reduced ELP in APC-EC, 57% less Lobo et al.

ur
and COD in EC: 0.079 kWh m−3 ENC in APC-EC than in DC-EC (2016)
real produced 99% TSS DC-EC/biochar:
Jo
water 99% turbidity 0.183 kWh m−3
5–14% COD
Al/Al CN‾, Ni, Cu, Trev= 60 s 80 A m−2 APC-EC: >20% – APC-EC: – – Improved REs and 6% higher ELC Pertile and
and Zn in higher CN‾, Ni, 36.29 g m−3 in APC-EC than in DC-EC Birriel (2017)
simulated and Cu removal DC-EC:
galvanic than DC-EC; 34.13 g m−3
effluent >98% Zn removal
in APC-EC and
DC-EC
Al/Al NaCl and Trev= 180 s 37.9 V – APC-EC: APC-EC: – – Effective de-passivation in APC- Wellner et al.
NaHCO3, in 2.56 kWh m−3 0.448 gr min−1 EC; 8% and 7% higher ENC and (2018)
synthetic WW DC-EC: DC-EC: ELC in APC-EC than in DC-EC
2.36 kWh m−3 0.419 gr min−1

29
Process performance
Opt. APC
Pollutant and
Anode/cathode waveform I, U, or j Removal Electrode Major outcomes References
solution Energy consumption Sludge production Operating
parameters efficiency consumption
(ENC) (SLP) cost (OC)
(RE) (ELC)
Al/Al Oil in Trev= 4 s 36 A m−2 APC-EC & DC- APC-EC: – – – Reduced ELP in APC-EC; Lower Bian et al.
simulated EC: 0.787-0.936 kWh m−3 ENC in APC-EC than in DC-EC (2019)
bilge water >99% oil DC-EC: after 24 h; >158% increase in ENC
0.378-0.977 kWh m−3 after 24 h in DC-EC
Al/Al Silica in real Trev= 20 s 160 A m−2 APC-EC & DC- APC-EC: APC-EC: – APC-EC: Soft powdery surface layer in APC- Yasri et al.
blowdown EC: 0.52 kWh m−3 156 g m−3 0.40 $ m−3 EC; Hard scale on the electrode in (2022)
water >95% SiO2 DC-EC; Lower ENC in APC-EC
than in DC-EC; 40% reduced RE
over time in DC-EC
Fe/Al Cr(VI) in Trev= 240 s 56–222 A APC-EC & DC- APC-EC: – – – Slower anode passivation and lower Keshmirizadeh

of
synthetic WW m−2 EC: 4-58 kWh m−3 ENC in APC-EC than in DC-EC et al. (2011)
98–99.9% Cr(VI)

ro
Al/Al, Fe/Fe*, Rhodamine 6G – 109 A m−2 APC-EC: APC-EC: APC-EC: – – Avoided ELP in APC-EC; Higher Zaleschi et al.
Fe/Al in synthetic 98% dye 6.84 kWh (kg dye)−1 1.73 kg Fe (kg RE and less ENC and ELC in APC- (2014)

-p
WW dye)−1 EC than in DC-EC; Fastest dye
removal with Fe/Fe

re
Al/Al*, Fe/Fe COD, color, Trev= 150 s 150 A m−2 APC-EC: APC-EC: APC-EC: APC-EC: APC-EC: Prevented ELP in APC-EC; 11% Eyvaz (2016)

lP
turbidity, TN, 91% COD 1.24 kWh (kg COD)−1 0.63 kg Al (kg 1.56 kg (kg COD)−1 2.80 $ m−3 higher COD removal and ~12%
and TP in real DC-EC: DC-EC: COD)−1 DC-EC: DC-EC: less ENC, ELC, SLP, and OC in
brewery 80% COD 1.41 kWh (kg COD)−1 DC-EC: 1.78 kg (kg COD)−1 3.18 $ m−3 APC-EC than in DC-EC

na
factory WW 0.72 kg Al (kg
COD)−1
Al/Fe*, Fe/Al Fluoride in
drinking water
Trev= 120 s
γ= 0.25
120 A m−2 APC-EC:
95% F‾ (20%
ur
APC-EC:
0.9 kWh m−3
APC-EC:
1.42 kg m−3
APC-EC:
0.108 kg m−3
– Avoided ELP and super FEs in
APC-EC
Alimohammadi
et al. (2019)
Jo
higher RE than
DC-EC)
Al/Fe, Fe/Al* Total and Trev= 120 s 90 A m−2 APC-EC: APC-EC: APC-EC: APC-EC: – APC-EC: 5% and 7% higher TH Yaghmaeian et
calcium γ= 0.25 98.26% TH 0.90 kWh m−3 1.03 kg m−3 0.098 kg m−3 and CH removal, 64% less ENC, al. (2020)
hardness in 87.69% CH DC-EC: DC-EC: DC-EC: 51% reduced ELC, and 10% lower
drinking water DC-EC: 2.47 kWh m−3 2.12 kg m−3 0.109 kg m−3 SLP than DC-EC
93.48% TH
80.23% CH
Al/Al, Fe/Fe*, Color and f= 50 Hz 40 A m−2 APC-EC: APC-EC: – APC-EC: lower SLP – Less ELP in APC-EC than in DC- Asaithambi et
Al/Fe, Fe/Al COD in real γ= 0.45 100% color 3.2 kWh m−3 than DC-EC EC; 9% higher REs and less ENC al. (2021)
distillery 95% COD DC-EC: in APC-EC than DC-EC; Higher
industrial DC-EC: 3.5 kWh m−3 COD removal and lower ENC with
effluent 90.57% color Fe/Fe
86.54% COD

30
Process performance
Opt. APC
Pollutant and
Anode/cathode waveform I, U, or j Removal Electrode Major outcomes References
solution Energy consumption Sludge production Operating
parameters efficiency consumption
(ENC) (SLP) cost (OC)
(RE) (ELC)
Fe/Fe Acid, basic, Trev= 300 s 2.73–27.32 A APC-EC: slightly APC-EC: APC-EC: 4% – – APC-EC: 9% and 4% less ENC and Secula et al.
and reactive m−2 enhanced REs 4.65 kWh (kg dye) −1 lower ELC than ELC than DC-EC (2013)
dyes in than DC-EC DC-EC: DC-EC
synthetic WW 5.12 kWh (kg dye) −1
Fe/Fe COD, TOC, Trev= 300 s 200 A m−2 H2O2/APC-EC: H2O2/APC-EC: H2O2/APC-EC: H2O2/APC-EC: H2O2/APC- Slower ELP in APC-EC than in Ozyonar and
phenol, CN‾, 92% COD 4.41 kWh m−3 1.08 kg m−3 1.89 kg m−3 EC: DC-EC; 35-42% higher REs in Karagozoglu
and SCN‾ in 90% TOC 5.64 € m−3 APC-EC than in DC-EC (2015)
raw coke WW H2O2/DC-EC:
50.2% COD
51% TOC

of
Fe/Fe COD, color, Trev= 3600 1.5 A APC-EC: – APC-EC: – – Reduced ELP in APC-EC; 137% Cesar Lopes
and turbidity s (PR 95% COD 0.2412 kg m−3 increase in ELC in APC-EC than Geraldino et al.

ro
in real dairy between 90% color DC-EC: DC-EC; Decreased REs over time (2016)
industry each exp.) 99% turbidity 0.1017 kg m−3 in DC-EC

-p
effluent

re
Fe/Fe COD and color Trev= 900 s 13 A m−2 APC-EC: APC-EC: – – – Reduced passive layer formation Asaithambi et
in real γ= 0.5 87% COD 1.31 kWh m−3 (lower and 20% higher COD removal in al. (2016)

lP
distillery 95% color ENC than DC-EC) APC-EC than in DC-EC
industrial DC-EC:
effluent 67% COD

na
Fe/Fe Zn2+, Cd2+, and Trev= 10 s 50–200 A APC-EC: higher APC-EC: lower ENC – – – The higher overall efficiency of Xu et al.
Mn2+ in real f= 5 kHz m−2 REs at 50 A m−2 at 50 A m−2 DC-EC than APC-EC. Greater REs (2018)
smelting WW γ= 0.4 DC-EC: higher
REs at 100-200 A ur
DC-EC: lower ENC at
150-200 A m−2
and lower ENC at lower j in APC-
EC
Jo
m−2
Fe/Fe Synthetic GW Trev= 1260, 8 A m−2 – APC-EC: APC-EC: – – Greater mass of the surface layer in Müller et al.
3150 s 0.21–0.22 kWh m−3 FE= 0.6–0.72 APC-EC than in DC-EC; 18-24% (2019)
DC-EC: DC-EC: less ENC and lower FEs in APC-
0.26 kWh m−3 FE= 0.75 EC than in DC-EC

Fe/Fe Cr(VI) in f= 5 kHz APC-EC: APC-EC: APC-EC: – APC-EC: – Reduced ELP in PDC-EC and Zhou et al.
synthetic WW γ= 0.3 +11, -3 A m−2 99.4% total Cr 760 kWh (kg 1.12 kg m−3 APC-EC; 80% less ENC and 90% (2020)
PDC-EC: PDC-EC: Cr(VI))−1 PDC-EC: more ENC in APC-EC than in DC-
25 A m−2 99.5% total Cr PDC-EC: 1.30 kg m−3 EC and PDC-EC, respectively; 40%
DC-EC: DC-EC: 400 kWh (kg DC-EC: and 14% lower SLP in APC-EC
20 A m−2 99.3% total Cr Cr(VI))−1 1.86 kg m−3 than in DC-EC and PDC-EC,
DC-EC: respectively
3800 kWh (kg
Cr(VI))−1

31
Process performance
Opt. APC
Pollutant and
Anode/cathode waveform I, U, or j Removal Electrode Major outcomes References
solution Energy consumption Sludge production Operating
parameters efficiency consumption
(ENC) (SLP) cost (OC)
(RE) (ELC)
Fe/Fe Cu2+ and COD f= 1 Hz 2 A m−2 Electro- Electro-Fenton/APC- – Electro- – Avoided ELP in APC-EC; 9% Xu et al.
in simulated Fenton/APC-EC: EC: Fenton/APC-EC: higher COD removal, 25% less (2021)
plating WW 99.59% Cu2+ 0.170 kWh m−3 0.928 kg m−3 ENC, and 9% less SLP in APC-EC
90.21% COD Electro-Fenton/DC- Electro-Fenton/DC- than in DC-EC
Electro- EC: EC:
Fenton/DC-EC: 0.226 kWh m−3 1.019 kg m−3
98.60% Cu2+
81.68% COD

Fe/Fe Polyvinyl f= 0.5 kHz 10 A m−2 APC-EC: APC-EC: APC-EC: APC-EC: – Reduced ELP in APC-EC; Stronger J. Zhang et al.
alcohol in γ= 0.3 99.73% PVA 0.049 kWh m−3 0.105 kg m−3 0.0674 g adsorption of flocs in APC-EC; 7% (2021)

of
synthetic WW DC-EC: DC-EC: DC-EC: DC-EC: higher RE and 90%, 45%, and 44%
92.82% PVA 0.480 kWh m−3 0.190 kg m−3 0.1193 g less ENC, ELC, and SLP in APC-

ro
EC than in DC-EC

-p
Fe, Ti Arsenic in Trev= 9 s 50 mA APC-EC: APC-EC: – – – Eliminated ELP in APC-EC; 40% Xin et al.
synthetic GW 98.4% As(tot) 0.101 kWh m−3 lower ENC in APC-EC than in DC- (2018)

re
DC-EC: DC-EC: EC
94.4% As(tot) 0.167 kWh m−3

lP
Fe/Ti Arsenic in Trev= 9 s: 40 mA APC-EC: APC-EC: – – – APC-EC: 85% higher RE than DC- Jiang et al.
synthetic GW 15 s 92% As(tot) 0.11 kWh m−3 EC (2018)

na
DC-EC:
7% As(tot)

Mg/Mg Indigo carmine


in synthetic
Trev= 450 s 50 A m−2 APC-EC: 43%
higher NPOC
ur
APC-EC:
22.1 kWh (kg COD)−1
– – – Avoided ELP and the formation of
highly branched flocs in APC-EC;
Donneys-
Victoria et al.
Jo
WW removal than DC- DC-EC: 55% higher ENC in APC-EC than (2019)
EC; Equal dye 14.3 kWh (kg COD)−1 in DC-EC; The formation of tightly
removal in APC- packed flocs in DC-EC
EC and DC-EC

Mg/Mg*, Indigo carmine Trev= 120 s 0.34 A APC-EC: – – – – APC-EC: 33% higher NPOC Donneys-
Mg/Al and Cl‾ in 90.4% NPOC removal and 61% less Cl− removal Victoria et al.
synthetic WW 7.8% Cl− than DC-EC (2020)
DC-EC:
57.1% NPOC
69.1% Cl−

32
Process performance
Opt. APC
Pollutant and
Anode/cathode waveform I, U, or j Removal Electrode Major outcomes References
solution Energy consumption Sludge production Operating
parameters efficiency consumption
(ENC) (SLP) cost (OC)
(RE) (ELC)
Al/Al*, Fe/Fe, Synthetic GW Trev= 600 s 10–130 A – APC-EC: higher ENC APC-EC: less FE – – APC-EC: ELP occurred for Fe/Fe; Chow and
Al/Ti-IrO2, m−2 than DC-EC for than DC-EC for Reduced ELP for Al/Al, Al/Ti-IrO2, Pham (2021)
Fe/Ti-IrO2 Fe/Fe; less ENC than Fe/Fe; and Fe/Ti-IrO2
DC-EC for Al/Al Comparable FE to
that of DC-EC for
Al/Al
MS/MS COD and oil Trev= 1.9 214 A m−2 APC-EC: APC-EC: APC-EC: APC-EC: increased APC-EC: Reduced ELP in APC-EC; Less Madhavan and
in Synthetic min 90% COD Higher ELC than DC- Lesser ENC than SLP compared to 0.65 $ (kg ENC and higher COD removal, FE, Antony (2021)
produced 88% oil EC DC-EC DC-EC COD)−1 and SLP in APC-EC than in DC-
water EC

of
St/St COD and color Trev= 5 min 144 A m−2 APC-EC: APC-EC: – APC-EC: – Reduced ELP in APC-EC; Reduced Sanei and

ro
in real leachate 35% COD 39.8 kWh (kg COD)−1 12 kg(TSS) (kg COD removal over time in DC-EC; Mokhtarani
50% color COD)−1 18% and 14% higher COD and (2022)

-p
color removal, and 12% and 51%
less ENC and SLP in APC-EC than

re
in DC-EC
* The best electrode combination
510

lP
I: Current; U: Voltage; j: Current Density; FE: Faradaic efficiency; WW: Wastewater; GW: Groundwater; MS: Mild steel; St: Steel
511

na
ur
Jo

33
512 Polarity reversal also allows a uniform distribution of the oxidation reactions between

513 electrodes, improving coagulant mass transfer and providing greater contact between

514 contaminants and coagulant agents in the solution (Pertile and Birriel, 2017; Gobbi et al., 2018;

515 Donneys-Victoria et al., 2019). It has been stated that the greater the distance between

516 electrodes, the longer the route the ions must take from one electrode to the other, and the

517 greater the likelihood that polarity may reverse while the coagulants are still in migration. This

518 increases their permanency and route, enhancing the likelihood of floc formation through

519 interaction with contaminants (Gobbi et al., 2018). Furthermore, flocs generated in APC-EC

f
oo
520 have been shown to have a more robust adsorption capacity than those produced in DC-EC

521

pr
(Zhou et al., 2020; Xu et al., 2021; J. Zhang et al., 2021). For example, Zhou et al. compared

the characteristics of flocs produced by APC-EC, PDC-EC, and DC-EC techniques and found
e-
522

523 that the flocs formed in APC-EC had a larger surface to adsorb Cr(VI) than those in DC-EC;
Pr

524 however, the flocs in PDC-EC have been reported to have a better spread performance than
al

525 those of others (Zhou et al., 2020).


rn

526 Several explanations have been proposed for the possible lower contaminant removal in
u
Jo

527 APC-EC compared to DC-EC, such as unsuccessful electrode de-passivation and inadequate

528 coagulant production, which can arise from improper selection of operating conditions,

529 including electrode material, current density, and waveform characteristics (e.g., PR time)

530 (Gelover-Santiago et al., 2012; Xu et al., 2018). For example, it has been asserted that at very

531 brief PR times, bubble gas generation from both the anode and cathode could be high and

532 instantaneous flotation could occur. As a result, coagulants could rise to the top of the EC cell

533 even before interacting with pollutants. In addition, enhanced bubble generation can impede

534 the transmission of electrons between electrodes, which can be detrimental to electrochemical

535 reactions and result in diminished pollutant removal (Madhavan and Antony, 2021).

34
536 Regarding energy efficiency, most authors demonstrated that reversing the polarity of the

537 electrodes reduced the voltage required compared to when a direct current was used (Lobo et

538 al., 2016; Xin et al., 2018; Yaghmaeian et al., 2020; Zhou et al., 2020; Asaithambi et al., 2021;

539 Madhavan and Antony, 2021; Xu et al., 2021; J. Zhang et al., 2021; Yasri et al., 2022). In the

540 case of the passivated electrode surface, passive overpotential—one of the elements

541 contributing to cell potential—will be greater. Periodic polarity reversal in APC-EC can lower

542 the passive overpotential and lead to diminished cell voltage, which eventually will reduce

543 energy utilization compared to DC-EC (Madhavan and Antony, 2021). Changing the current

f
oo
544 waveform from DC to APC, for example, has been proven to reduce energy consumption by

545

pr
up to 90% (J. Zhang et al., 2021). Of course, in this study, in addition to passivation reduction,

the current breaks during original and reversed polarity also helped to reduce energy
e-
546

547 consumption significantly. On the contrary, the interruption of the circuit caused by the reversal
Pr

548 cycles of the polarity might increase the cell voltage for a few seconds afterward, resulting in
al

549 higher energy consumption (Bian et al., 2019; Donneys-Victoria et al., 2019). Bian et al.
rn

550 reported that the energy usage under APC power was higher than under DC power at the
u

551 beginning of the operation due to the higher voltage at the start of every shift of the electric
Jo

552 field direction (Bian et al., 2019). This research group (Bian et al., 2019) reported that the

553 advantage of APC power became apparent after 24 hours of operation when APC-EC

554 consumed less power than DC-EC, probably owing to reduced passivation.

555 The literature includes contradictory findings regarding electrode consumption. Some

556 authors reported that diminished passivation in APC-EC led to increased current density and

557 higher electrode consumption (and thus higher faradaic efficiency), while in DC-EC, the

558 amount of material lost from the electrodes was smaller due to passivation (Pi et al., 2014;

559 Cesar Lopes Geraldino et al., 2016; Pertile and Birriel, 2017; Wellner et al., 2018). Based on

560 the literature, electrode usage has often been higher in APC-EC than in DC-EC when Al

35
561 electrodes are used. At prolonged PR times, acidic/basic boundary layers around the Al

562 electrode have adequate time to develop after polarity reversal, which contributes to the

563 chemical corrosion of Al electrodes and de-passivation (Fuladpanjeh‐Hojaghan et al., 2019;

564 Yasri et al., 2022). Another explanation for higher electrode consumption in APC-EC could be

565 the lower possibility of side reactions caused by polarity reversal when the Al electrode is used

566 since aluminum is less reactive than iron within water's electrochemical stability range

567 (Roberge, 2008). In contrast, some studies indicated that fewer electrodes were consumed in

568 APC-EC than in DC-EC (Secula et al., 2013; Zaleschi et al., 2014; Eyvaz, 2016; Yaghmaeian

f
oo
569 et al., 2020; J. Zhang et al., 2021). This behavior can be attributed to uniform electrode

570

pr
corrosion in APC-EC caused by cyclical energization between anode and cathode, ensuring

longer electrode life (Eyvaz, 2016; J. Zhang et al., 2021). Also, shorter actual electrification
e-
571

572 times in APC-EC can reduce electrode consumption (Yaghmaeian et al., 2020). Moreover,
Pr

573 unsuccessful electrode de-passivation and the occurrence of electrochemical side reactions in
al

574 APC-EC can result in decreased electrode consumption and coagulant production (Müller et
rn

575 al., 2019; Chow and Pham, 2021).


u
Jo

576 According to almost all the studies reviewed, the amount of sludge generated during

577 APC-EC was less than that produced during DC-EC (Eyvaz, 2016; Yaghmaeian et al., 2020;

578 Zhou et al., 2020; Asaithambi et al., 2021; Xu et al., 2021; J. Zhang et al., 2021; Sanei and

579 Mokhtarani, 2022). For instance, Zhang et al. indicated that sludge production under APC

580 conditions was about 44% less than in DC (J. Zhang et al., 2021). The reason for this could be

581 the shorter electrification time in APC-EC compared to DC-EC, which results in less electrode

582 usage and sludge generation. As another example, it was discovered that APC-EC produced

583 about 40% and 14% less sludge than PDC-EC and DC-EC, respectively, when treating Cr(VI)

584 from synthetic wastewater (Zhou et al., 2020). In contrast, it was reported that reversing the

585 electrode polarity enhanced sludge production owing to a significantly improved electrode

36
586 corrosion and coagulation process (Madhavan and Antony, 2021). Finally, since the overall

587 energy usage, electrode consumption, and sludge generation in most studies were lower in the

588 case of APC-EC, the operating cost for this technique is expected to be less compared to that

589 of DC-EC, as proved in some studies (Eyvaz et al., 2009; Eyvaz, 2016).

590

591 7. Sinusoidal alternating current electrocoagulation (AC-EC)

592 Several studies have operated EC with sinusoidal alternating current (AC-EC) to reduce

593 passivation and improve efficiency. The anode and cathode in AC-EC are switched periodically

f
oo
594 based on the sinusoidal waveform; thus, each electrode is potentially an exchangeable anode

595
pr
or cathode. An overview of studies that evaluated the passivation mitigation capability of AC-
e-
596 EC and compared its performance with DC-EC is presented in Table 3. As indicated in Table
Pr

597 3, AC-EC has been mainly operated at 50 Hz and evaluated with different electrode materials

598 for various applications. Most authors have reported that AC-EC and DC-EC exhibit
al

599 comparable pollutant removal, with AC-EC having a slight advantage (Vasudevan and
rn

600 Lakshmi, 2012; Kamaraj et al., 2013; Karamati-Niaragh et al., 2019; Kong et al., 2020; Xu et
u

601 al., 2020). Higher removal efficiencies in AC-EC could probably be due to the occasional shift
Jo

602 in the anode and cathode operations, which reduces the load on the function of each electrode,

603 postpones passivation, and improves metal (oxy)hydroxide formation (Alkhatib et al., 2020).

604 Moreover, similar to APC-EC, the metallic cations scatter more uniformly throughout the cell

605 in AC-EC due to the corrosion of both the anode and cathode, which improves coagulant mass

606 transfer and facilitates the interactions of coagulants with the pollutant, thereby enhancing

607 efficiency. It was also found that the flocs produced by AC-EC had many pores, a loose

608 structure, and a large specific surface area, which was favorable for the adsorption of heavy

609 metals, while the flocs formed by DC-EC had a compact structure. As such, the Cr(VI) content

610 adsorbed by Fe(OH)3 flocs in AC-EC was more significant than in DC-EC (Xu et al., 2019).

37
611 Table 3. Summary of reported studies on the comparison of AC-EC with DC-EC technique

Process performance
Pollutant and AC
Anode/cathode I, U, or j Sludge Major outcomes Reference
solution frequency Pollutant removal Energy consumption Electrode consumption Operation
production
efficiency (RE) (ENC) (ELC) cost (OC)
(SLP)
Al/Al F‾ in deionized 50 Hz 100 A AC-EC: AC-EC: – – – Avoided ELP and uniform Al corrosion in Vasudevan
water m−2 93.0% F‾ 1.88 kWh m−3 AC-EC; Slightly higher RE and 26% less et al. (2011a)
DC-EC: DC-EC: ENC in AC-EC than in DC-EC
91.5% F‾ 2.54 kWh m−3
Al/Al Cd in synthetic 50 Hz 20 A m−2 AC-EC: AC-EC: – – – Slightly higher RE and 55% less ENC in Vasudevan
WW 97.5% Cd 0.45 kWh m−3 AC-EC than DC-EC et al. (2011b)
DC-EC: DC-EC:
96.2% Cd 1.00 kWh m−3

of
Al/Al Oil, color, and AC-EC: 3A AC-EC: AC-EC: AC-EC: – – Reduced ELP in AC-EC; Similar RE and Cerqueira et

ro
turbidity in 60 Hz 94.3% oil 0.28 kWh m−3 0.12 kg m−3 ENC in AC-EC and APC-EC; 33% less al. (2014)
simulated APC-EC: APC-EC: APC-EC: APC-EC: ELC in AC-EC than in APC-EC

-p
produced water Trev= 5 94.1% oil 0.28 kWh m−3 0.18 kg m−3
min

re
Al/Al NO3‾ in tap water 50 Hz 70–80 A AC-EC: Slightly higher ENC AC-EC: – AC-EC: Minimized ELP and uniform Al corrosion in Karamati-
m−2 76.67% NO3‾ in AC-EC than DC- 4.96 kg Al (kg NO3‾)−1 29 $ (kg AC-EC; 75% reduced ELC and 46% lower Niaragh et

lP
DC-EC: EC. DC-EC: NO3‾)−1 OC in AC-EC than in DC-EC for similar RE al. (2019)
74.67% NO3‾ 19.96 kg Al (kg DC-EC:

na
NO3‾)−1 54 $ (kg
NO3‾)−1
Al/Al COD and total 50 Hz 43 A m−2 AC-EC: – AC-EC: – – Decreased ELP in AC-EC; 23.5% and Alkhatib et
phosphorus in
secondary treated
66.8% COD
68.2% TP
ur 0.15 g
DC-EC:
22.2% higher COD and TP removal, and
93% less ELC in AC-EC than DC-EC
al. (2020)
Jo
WW DC-EC: 2.12 g
43.3% COD
46% TP
Al/Al Nickel in 50 Hz 50–90 A Almost equal AC-EC: AC-EC: – – Uniform Al corrosion in AC-EC; 16% and Arabameri et
synthetic WW m−2 nickel removal in 29.2 kWh (kg Ni)−1 1.20 kg Al (kg Ni)−1 48% lower ENC and ELC in AC-EC than in al. (2022)
AC-EC and DC- DC-EC: DC-EC: DC-EC for similar RE
EC 34.9 kWh (kg Ni)−1 2.3 kg Al (kg Ni)−1

Al/Al*, Fe/Fe COD, color, and 60 Hz 18.5– AC-EC & DC- – Almost double ELC in – – Higher COD removal with Al than with Fe Rodrigues
turbidity in 55.5 A EC: DC-EC than AC-EC. in both techniques; 50% less ELC in AC-EC Pires da
synthetic oil-in- m−2 >92% COD than in DC-EC for similar REs Silva et al.
water emulsion >96% color (2015)
>99.5% turbidity

38
Process performance
Pollutant and AC
Anode/cathode I, U, or j Sludge Major outcomes Reference
solution frequency Pollutant removal Energy consumption Electrode consumption Operation
production
efficiency (RE) (ENC) (ELC) cost (OC)
(SLP)
Al/Al*, Fe/Fe F‾ in distilled – 40 V AC-EC: – – – – Higher REs with Al than with Fe; 5.75% Ghanizadeh
water 81.75% F‾ less RE in AC-EC than in DC-EC et al. (2016)
DC-EC:
87.50% F‾
Al/Al, Fe/Fe* Acid Red 18 in 50 Hz 30–130 AC-EC: Higher ENC in AC- AC-EC: Lower SLP in AC-EC: Reduced ELP and uniform electrode Payami
synthetic WW A m−2 71% dye (Fe) EC for Al; Higher 0.15 g (Fe) AC-EC than 9.4 $ (kg corrosion in AC-EC; 14% less RE (with Fe), Shabestar et
42% dye (Al) ENC in DC-EC for 0.25 g (Al) DC-EC dye)−1 92% and 64% lower Fe and Al usage, and al. (2021)
DC-EC: Fe. DC-EC: DC-EC: 36% lower OC in AC-EC than in DC-EC;
85% dye (Fe) 1.78 g (Fe) 14.6 $ (kg Higher efficiency of AC-EC with Fe
42% dye (Al) 0.70 g (Al) dye)−1
20 A m−2 – – –

of
Fe/Fe Cd in synthetic 50 Hz AC-EC: AC-EC: Uniform Fe corrosion in AC-EC; Slightly Vasudevan
WW 98.1% Cd 0.734 kWh m−3 higher RE and 48% lower ENC in AC-EC and Lakshmi

ro
DC-EC: DC-EC: than in DC-EC (2012)
97.3% Cd 1.413 kWh m−3

-p
Fe/Fe Cr(VI) in AC-EC: 2.7 A m−2 AC-EC, PDC-EC, AC-EC: AC-EC: – – Reduced ELP in AC-EC than in DC-EC and Xu et al.
synthetic WW 50 Hz DC-EC: 0.0267 kW h−1 0.0128 g PDC-EC; Better adsorption of flocs in AC- (2019)

re
PDC-EC: >99% Cr(VI) PDC-EC: DC-EC: EC; 13% and 6% less ENC in AC-EC than
f= 1 kHz, (removal rate: 0.0284 kW h−1 0.0149 g in DC-EC and PDC-EC; 14% less ELC in

lP
γ= 0.2 AC>PDC>DC) DC-EC: AC-EC than in DC-EC; Reduced FE with
0.0307 kW h−1 time in DC-EC

na
Fe/Fe COD in 50 Hz 1000 A Fenton/AC-EC: – – Lower SLP in – Fenton/AC-EC: Reduced ELP, slightly Kong et al.
simulated m−3 91.0% COD Fenton/AC-EC higher RE, a larger specific area of flocs, (2020)

ur
electroplating Fenton/DC-EC: than in and less SLP than Fenton/DC-EC
WW 87.25% COD Fenton/DC-EC
Jo
Fe/Fe Cu2+ in synthetic 50 Hz 0.25 A AC-EC: AC-EC: AC-EC: AC-EC: – Prevented ELP and enhanced adsorption of Xu et al.
and real printed m−2 99.2% Cu2+ 0.99 kWh m−3 12 g m−3 18.9 g m‾3 Cu2+ in AC-EC; Slightly higher RE, 26%, (2020)
circuit WW DC-EC: DC-EC: DC-EC: DC-EC: 45%, and 54% lower ENC, ELC, and SLP
97.5% Cu2+ 1.342 kWh m−3 22 g m−3 41.2 g m‾3 in AC-EC than in DC-EC
Fe-Fe Total phosphorus 50 Hz 21.2 A AC-EC: – – – – Increased specific area and better adsorption Zhou et al.
in synthetic WW m−2 90.9% TP (higher of flocs in AC-EC than in DC-EC (2021)
RE than DC-EC)
Fe/Fe*, SS/SS Pb and Zn in real – 60 A m−2 AC-EC: AC-EC: AC-EC: AC-EC: – Slightly lower REs and 65%, 39%, and 8% Mansoorian
battery industry 96.7% Pb 0.69 kWh m−3 0.69–0.72 kg Fe m−3 0.084 kg m‾3 less ENC, ELC, and SLP in AC-EC than in et al. (2014)
WW 95.2% Zn DC-EC: DC-EC: DC-EC: DC-EC; Higher efficiencies with Fe than
DC-EC: 1.97 kWh m−3 1.13–1.17 kg Fe m−3 0.091 kg m‾3 with SS
97.2% Pb
95.5% Zn

39
Process performance
Pollutant and AC
Anode/cathode I, U, or j Sludge Major outcomes Reference
solution frequency Pollutant removal Energy consumption Electrode consumption Operation
production
efficiency (RE) (ENC) (ELC) cost (OC)
(SLP)
Fe/Fe*, SS/SS Malathion in – 100 A AC-EC: AC-EC: AC-EC: AC-EC: – Similar RE and SLP in both techniques; Khaghani et
synthetic WW m−2 95.6% malathion 0.91 kWh m−3 1.06 kg Fe m−3 0.093 kg m‾3 71% and 42% less ENC and ELC in AC-EC al. (2020)
DC-EC: DC-EC: DC-EC: DC-EC: than in DC-EC
95.8% malathion 3.12 kWh m−3 1.82 kg Fe m−3 0.095 kg m‾3
Zn/Zn Cd in synthetic 50 Hz 20 A m−2 AC-EC: AC-EC: – – – Reduced ELP and uniform Zn corrosion in Vasudevan
WW 97.8% Cd 0.665 kWh m−3 AC-EC; 46% lower ENC in AC-EC than in and Lakshmi
DC-EC: DC-EC: DC-EC for almost equal RE (2011)
96.9% Cd 1.236 kWh m−3

Zn/Zn Fe in water 50 Hz 6 A m−2 AC-EC: AC-EC: – – – Diminished ELP and uniform Zn corrosion Vasudevan
99.6% Fe 0.625 kWh m−3 in AC-EC; 37% less ENC in AC-EC than in (2012)

of
DC-EC: DC-EC: DC-EC for almost equal RE
99.1% Fe 0.991 kWh m−3

ro
Mg/Mg Arsenate in water 50 Hz 20 A m−2 AC-EC: AC-EC: – – – AC-EC: 30% less ENC than DC-EC for Vasudevan
98.3% arsenate 0.724 kWh m−3 almost equal RE. et al. (2012)

-p
DC-EC: DC-EC:
1.035 kWh m−3

re
97.9% arsenate

Mg/Mg Copper in water 50 Hz 2.5 A m−2 AC-EC: AC-EC: – – – ELP in AC-EC in the presence of carbonate; Kamaraj et

lP
97.8% Cu 0.634 kWh m−3 Uniform Mg corrosion in AC-EC; 36% less al. (2013)
DC-EC: DC-EC: ENC in AC-EC than in DC-EC for almost
0.996 kWh m−3

na
97.2% Cu equal RE

Pb/Pb Indigo Carmine – 0.5 A m−2 AC-EC: AC-EC: – – – Ductile and frangible surface layers in AC- Othmani et

ur
and Methylene 100% IC 0.44×10-6 kWh m−3 EC and DC-EC, respectively; Uniform al. (2017)
Blue in synthetic 99% MB (IC) electrode corrosion in AC-EC; Slightly
Jo
WW DC-EC: 4.89×10-6 kWh m−3 higher RE and 36–78% less ENC in AC-EC
98.8% IC (MB) than in DC-EC
96.5% MB DC-EC:
2.03×10-6 kWh m−3
(IC)
7.61×10-6 kWh m−3
(MB)
* The best electrode combination
612
I: Current; U: Voltage; j: Current Density; WW: Wastewater; ELP: Electrode passivation; SS: Stainless steel
613

40
614 However, in some studies, AC-EC was less efficient than DC-EC in removing pollutants

615 (Mansoorian et al., 2014; Ghanizadeh et al., 2016; Payami Shabestar et al., 2021). For example,

616 the findings of our research group (Payami Shabestar et al., 2021) showed that DC-EC was

617 more efficient than AC-EC in removing Acid Red 18 when using Fe electrodes, but the removal

618 efficiencies for Al electrodes were nearly equal in both techniques. This is because the higher

619 final pH of DC-EC was more conducive to the formation of Fe-(oxy)hydroxides, resulting in a

620 higher pollutant removal efficiency (Payami Shabestar et al., 2021). Besides, the rate and

621 mechanism of electrode passivation depend on the electrode material, so various electrode

f
oo
622 materials passivate differently (Chow and Pham, 2021; Payami Shabestar et al., 2021). As a

623

pr
result, employing an AC waveform to prevent electrode passivation could have different results

based on the electrode material. Furthermore, based on the literature, the improvements in
e-
624

625 pollutant removal efficiency in AC-EC compared to DC-EC are less significant than the
Pr

626 differences between APC-EC and DC-EC (see Tables 2 and 3). The PR time in AC-EC (often
al

627 20 ms PR time, i.e., 50 Hz PR frequency) might be shorter than the time required for charging
rn

628 the electrode double-layer, making this technique inefficient in some experimental conditions,
u

629 especially when treating solutions with poor conductivity (Ingelsson et al., 2020). Lower
Jo

630 faradaic efficiencies could also contribute to decreased contaminant removal in AC-EC than in

631 DC-EC (Mansoorian et al., 2014).

632 From the viewpoint of energy usage, most research indicates that AC-EC is more

633 efficient than DC-EC. Based on the data reported in Table 3, AC-EC consumed approximately

634 13–80% less energy than DC-EC (Vasudevan et al., 2011a; Vasudevan and Lakshmi, 2012;

635 Kamaraj et al., 2013; Mansoorian et al., 2014; Othmani et al., 2017; Xu et al., 2019, 2020;

636 Khaghani et al., 2020; Arabameri et al., 2022). The passivating layer in DC-EC prevents

637 electrode corrosion, which increases passivation overpotential and energy consumption.

638 However, alternating the polarity between the anode and cathode in AC-EC reduces the

41
639 passivating film growth (Xu et al., 2020). In contrast, a few studies have reported that the

640 energy consumption in AC-EC tended to be greater than that of DC-EC, but the differences

641 were not significant (Karamati-Niaragh et al., 2019; Payami Shabestar et al., 2021). For

642 instance, in the study accomplished by our team (Karamati-Niaragh et al., 2019) to eliminate

643 nitrate from water, DC-EC was slightly more energy-efficient than AC-EC at current intensities

644 between 2 and 4 A. The higher energy consumption in AC-EC may be related to the types of

645 contaminants and their complexes of removal mechanisms.

All the reviewed studies reported that using the AC waveform instead of the DC

f
646

oo
647 waveform lowers electrode consumption. It was shown that electrode utilization in AC-EC can

648
pr
be reduced by about 14–93% compared to DC-EC, regardless of the electrode material used
e-
649 (Mansoorian et al., 2014; Rodrigues Pires da Silva et al., 2015; Karamati-Niaragh et al., 2019;
Pr

650 Xu et al., 2019, 2020; Alkhatib et al., 2020; Khaghani et al., 2020; Payami Shabestar et al.,

651 2021; Arabameri et al., 2022). Through scanning electron microscopy (SEM) imaging, it was
al

652 shown that electrode oxidation in DC-EC occurs at the same preferential spots of the electrode
rn

653 surface due to the unidirectional current. Therefore, the electrode corrodes irregularly and non-
u
Jo

654 uniformly (Karamati-Niaragh et al., 2019). Conversely, the returning current in AC-EC forms

655 new oxidation spots, which decreases the locally attacked portion of the anode/cathode surface

656 and extends the uniformly attacked area (Payami Shabestar et al., 2021). Thus, electrodes in

657 DC-EC are more susceptible to localized corrosion and precipitation of minerals than those in

658 AC-EC. As a result, the corrosion type changes from pitting corrosion in DC-EC to uniform

659 corrosion in AC-EC, decreasing the electrode mass loss.

660 Furthermore, frequent changes in electrode polarity in AC-EC increase the likelihood of

661 side reactions, which divert the current away from electrode corrosion. For example, it has been

662 reported that the Pb2+ and Zn2+ ions in battery industry wastewater could be reduced at the

663 cathode and Pb2+ oxidized to Pb4+ at the anode repeatedly following each polarity reversal,

42
664 diverting the current away from electrode corrosion and lowering coagulant generation

665 (Mansoorian et al., 2014; Nidheesh et al., 2022). In the reviewed studies, the AC-EC has

666 utilized fewer Al electrodes, in contrast to the APC-EC, which often has been demonstrated to

667 have more Al electrode mass loss than the DC-EC. With the frequent polarity reversal in AC-

668 EC, the pH boundary layers change repeatedly and have less time to reconstitute. As a result,

669 low acidic/basic pH regions develop at the electrode interfaces, which lessens the likelihood of

670 Al electrodes' chemical corrosion (Yasri et al., 2022). For example, Cerqueira et al. showed

671 that AC-EC (60 Hz) consumed about 50% less aluminum mass than APC-EC (a 5 min PR

f
oo
672 time) for the same oil and grease removal (Cerqueira et al., 2014). Xu et al. compared the AC-

EC (50 Hz), PDC-EC (1 kHz and γ= 0.2), and DC-EC using Fe electrodes and found that the
673

pr
AC-EC was the best and most energy-efficient technique in Cr(VI) removal; however, it had a
e-
674

675 lower faradaic efficiency than the DC-EC (Xu et al., 2019). The lower faradaic efficiency in
Pr

676 AC-EC was attributed to Fe2+ species reduction to Fe(0) because Fe electrodes instantly turned
al

677 into a cathode after corroding as an anode (Xu et al., 2019).


rn

678 EC studies have shown a decrease in sludge production when using the AC waveform
u
Jo

679 instead of the DC waveform due to the uniform dissolution of anode and cathode and decreased

680 electrode consumption (Mansoorian et al., 2014; Kong et al., 2020; Xu et al., 2020; Payami

681 Shabestar et al., 2021). For example, it has been reported that sludge production can be

682 decreased by up to 54% by altering the current waveform from DC to AC (Xu et al., 2020).

683 Therefore, since EC's operating costs mainly include the cost of energy, electrode usage, and

684 sludge disposal, AC-EC is more cost-effective than DC-EC. The previous work of our research

685 group (Karamati-Niaragh et al., 2019) indicated that using an AC waveform instead of DC in

686 the EC process could even drop the operating cost by about half.

687

43
688 8. The suitable current waveforms for treating various solution matrices

689 Solution chemistry is a critical factor that should be considered while selecting the proper

690 current waveform in the EC process. Based on Fig. SM-2, various water and wastewater

691 matrices have been treated in EC under PDC, APC, and AC waveforms that can be categorized

692 as follows: water, groundwater, synthetic industrial wastewater, real industrial wastewater, and

693 municipal wastewater. For water treatment, the AC has been the most widely used waveform

694 that has exhibited comparable pollutant removal (e.g., fluoride, iron, arsenate, and copper) to

695 the DC while requiring significantly less energy (Vasudevan et al., 2011a, 2012; Vasudevan,

f
oo
696 2012; Kamaraj et al., 2013). In addition, the APC waveform has shown remarkable

697
pr
performance in removing fluoride (Alimohammadi et al., 2019) and hardness (Yaghmaeian et
e-
698 al., 2020) from drinking water. However, only the APC waveform has been used for treating
Pr

699 groundwater, and conflicting results have been reported. It was found that 21 and 52.5 min PR

700 times were insufficient to successfully de-passivate Fe electrodes (Müller et al., 2019), while a
al

701 30 min PR time was enough to prevent Al passivation when treating groundwater (Mohora et
rn

702 al., 2014). These conflicting results can be attributed to the different operational conditions,
u

703 such as the type and concentration of ionic species in groundwater and the electrode material.
Jo

704 For instance, it was found that Al electrodes were not covered by Ca and Mg minerals to the

705 same extent as the Fe electrode when treating groundwater by PR-EC due to the chemical

706 corrosion of aluminum, which contributes to de-passivation (Chow and Pham, 2021).

707 Synthetic industrial wastewaters containing pollutants, such as oil, COD, dyes, and heavy

708 metals (Pi et al., 2014; Bian et al., 2019; Zhou et al., 2021; Arabameri et al., 2022), have been

709 the most common matrices treated in EC under various current waveforms. According to the

710 literature, the APC and PDC have provided a better balance of pollutant removal, energy

711 consumption, electrode usage, and sludge production (Eyvaz et al., 2009; Ren et al., 2011;

712 Zhou et al., 2020; J. Zhang et al., 2021). In most cases, the combination of the APC waveform

44
713 and Al electrodes resulted in superior de-passivation capability when treating synthetic

714 industrial wastewaters (Pi et al., 2014; Wellner et al., 2018). Meanwhile, the AC has shown

715 almost similar contaminant removal compared to the DC (Karamati-Niaragh et al., 2019; Xu

716 et al., 2019; Arabameri et al., 2022). Based on Fig. SM-2, researchers have been more inclined

717 to operate EC with the APC and PDC waveforms when treating real industrial wastewater. The

718 majority of research has shown that applying the APC waveform in EC results in improved

719 treatment efficiency and diminished energy and electrode usage (Eyvaz, 2016; Asaithambi et

720 al., 2021; Sanei and Mokhtarani, 2022). However, due to the complex chemistry of real

f
oo
721 industrial wastewaters and the presence of a high concentration of interfering ions, their

722

pr
treatment could be more challenging compared to other water and wastewater matrices

(Mansoorian et al., 2014; Xu et al., 2018; Rodrigues et al., 2020). Besides, the PDC waveform
e-
723

724 has received the most attention for treating municipal wastewater and has been proven to
Pr

725 outperform the DC in removing COD from municipal wastewater while being more energy-
al

726 efficient (H. Nguyen, 2020; Que et al., 2021). However, low pharmaceutical and antibiotic
rn

727 removal from municipal wastewater has been reported using the PDC waveform (Ensano et
u

728 al., 2019; Oliveira et al., 2021).


Jo

729 Accordingly, various water and wastewater matrices can be efficiently treated in EC

730 under the PDC, APC, and AC waveforms. However, when selecting the proper current

731 waveform, it is crucial to carefully weigh the trade-offs between de-passivation, pollutant

732 removal efficiency, energy and electrode use, sludge formation, and operating cost. Based on

733 the literature, a few studies have compared the efficacy of PDC-EC, APC-EC, and AC-EC in

734 treating different water and wastewater matrices. Thus, there is still a need for more research

735 on choosing the current waveform and waveform characteristics for various combinations of

736 electrode material and water and wastewater matrices. It is also suggested to compare the

737 efficacy of various current waveforms in treating synthetic and real wastewater.

45
738

739 9. Conclusions

740 Despite various advantages, the EC process is subjected to high energy and electrode

741 consumption and electrode passivation, which prevent it from being fully commercialized and

742 practically used. To overcome these shortcomings, the role of the current waveform in the EC

743 process has received much consideration, and it has been the topic of intensive research over

744 the past two decades. This review paper outlines the PDC, APC, and AC waveforms employed

745 in the EC process as alternatives to DC. The role of these waveforms in de-passivation was

f
oo
746 thoroughly explored, and their influence on the overall EC performance was compared to DC.

747
pr
Besides, the impact of current waveform parameters on PDC-EC and APC-EC was evaluated.
e-
748 Moreover, the suitable current waveforms for treating various water and wastewater matrices
Pr

749 were discussed. Based on this review, the PDC-EC, APC-EC, and AC-EC techniques have

750 often exhibited similar or greater pollutant removal than DC-EC while mitigating passivation,
al

751 consuming less energy and electrode, producing lower sludge, and decreasing operational
rn

752 costs. They could also provide other benefits, such as uniform corrosion of electrodes,
u

753 enhanced coagulant mass transport, and better adsorption properties of flocs.
Jo

754

755 10. Future outlooks

756 Based on this review, the following are the main aspects of future research on mitigating

757 electrode passivation and enhancing EC efficiency using PDC, APC, and AC waveforms:

758 (1) EC operation under PDC, APC, and AC waveforms has been demonstrated in most

759 studies to offer a better balance of pollutant removal, energy requirements, electrode

760 usage, sludge production, and operating expenses compared with conventional DC.

761 This improvement is often attributed to reduced passivation. However, few

762 researchers have analyzed the electrode surface to support this hypothesis. More

46
763 research on this subject will help clarify the impact of various current waveforms on

764 the properties of the electrode surface layer.

765 (2) Some authors have reported that the application of PDC, APC, and AC waveforms

766 in the EC process, in addition to de-passivating electrodes, could contribute to

767 uniform corrosion of electrodes, enhance coagulant mass transfer, and improve the

768 adsorption properties of flocs. These mechanisms deserve further attention, and more

769 research in this area is encouraged.

770 (3) There have been conflicting findings on how current waveform parameters affect the

f
oo
771 EC process and the optimal amounts of these parameters to select. The optimal pulse

772

pr
frequency in PDC-EC and the suggested PR time and pulse frequency values in APC-

EC differed widely. Besides, no studies have explored the impact of current


e-
773

774 frequency on AC-EC. Thus, more research is needed regarding optimizing waveform
Pr

775 parameters in all the above-mentioned techniques.


al

776 (4) Based on the solution chemistry, there would probably be an optimal current
rn

777 waveform and electrode material. Hence, more research on selecting the current
u

778 waveform (DC, PDC, APC, and AC), waveform characteristics (e.g., PR time, pulse
Jo

779 frequency, and duty cycle), and electrode material for various water and wastewater

780 matrices is required.

781 (5) Most studies evaluated the impact of current waveforms on the EC in small-scale

782 batch reactors, while most commercial applications require continuous flow mode.

783 Thus, future research should also consider the long-term continuous flow modes of

784 PDC-EC, APC-EC, and AC-EC techniques in pilot-scale systems to simulate

785 industrial implementation.

786

47
787 CRediT authorship contribution statement

788 Javad Abdollahi: Conceptualization, Investigation, Literature review, Writing- original draft

789 preparation, Writing- reviewing and editing, Visualization. Mohammad Reza Alavi

790 Moghaddam: Conceptualization, Writing- reviewing and editing, Supervision. Sajjad

791 Habibzadeh: Conceptualization, Writing- reviewing and editing, Supervision.

792

793 Declaration of Competing Interest

794 The authors declare that they have no known competing financial interests or personal

f
oo
795 relationships that could have appeared to influence the work reported in this paper.

796
pr
e-
797 Acknowledgments
Pr

798 The authors gratefully acknowledge the financial support of the Amirkabir University of

799 Technology (Tehran Polytechnic).


al

800
rn

801 References
u
Jo

802 Akhbarati, R., Keshmirizadeh, E., Modarress, H., 2017. Uptake of acid blue 113 dye from
803 aqueous solution by sludge/floc nanoparticles in electrocoagulation process. Desalin.
804 Water Treat. 87, 314–325. https://doi.org/10.5004/dwt.2017.21294
805 Alimohammadi, M., Mesdaghinia, A., Shayesteh, M.H., Mansoorian, H.J., Khanjani, N.,
806 2019. The efficiency of the electrocoagulation process in reducing fluoride: application
807 of inductive alternating current and polarity inverter. Int. J. Environ. Sci. Technol. 16,
808 8239–8254. https://doi.org/10.1007/s13762-019-02297-4
809 AlJaberi, F.Y., Alardhi, S.M., Ahmed, S.A., Salman, A.D., Juzsakova, T., Cretescu, I., Le, P.-
810 C., Chung, W.J., Chang, S.W., Nguyen, D.D., 2022. Can electrocoagulation technology
811 be integrated with wastewater treatment systems to improve treatment efficiency?
812 Environ. Res. 214, 113890. https://doi.org/10.1016/j.envres.2022.113890
813 Alkhatib, A.M., Hawari, A.H., Hafiz, Mhd.A., Benamor, A., 2020. A novel cylindrical
814 electrode configuration for inducing dielectrophoretic forces during electrocoagulation.
815 J. Water Process Eng. 35, 101195. https://doi.org/10.1016/j.jwpe.2020.101195

48
816 Al-Raad, A.A., Hanafiah, M.M., 2021. Removal of inorganic pollutants using
817 electrocoagulation technology: A review of emerging applications and mechanisms. J.
818 Environ. Manag. 300, 113696. https://doi.org/10.1016/j.jenvman.2021.113696
819 Amrose, S.E., Bandaru, S.R.S., Delaire, C., van Genuchten, C.M., Dutta, A., DebSarkar, A.,
820 Orr, C., Roy, J., Das, A., Gadgil, A.J., 2014. Electro-chemical arsenic remediation: Field
821 trials in West Bengal. Sci. Total Environ. 488–489, 539–546.
822 https://doi.org/10.1016/j.scitotenv.2013.11.074
823 An, B.-H., Xu, D.-M., Geng, R., Cheng, Y., Qian, R.-B., Tang, X.-C., Fan, Z.-Q., Chen, H.-
824 B., 2022. The treatment effects of various target pollutant in real coal gasification gray
825 water by coupling pulse electrocoagulation with chemical precipitation methods.
826 Chemosphere 136898. https://doi.org/10.1016/j.chemosphere.2022.136898
827 Arabameri, A., Alavi Moghaddam, M.R., Azadmehr, A.R., Payami Shabestar, M., 2022. Less

f
828 energy and material consumption in an electrocoagulation system using AC waveform

oo
829 instead of DC for nickel removal: Process optimization through RSM. Chem. Eng.
830 Process.: Process Intensif. 174, 108869. https://doi.org/10.1016/j.cep.2022.108869
831
832 pr
Arroyo, M.G., Pérez-Herranz, V., Montañés, M.T., García-Antón, J., Guiñón, J.L., 2009.
Effect of pH and chloride concentration on the removal of hexavalent chromium in a
e-
833 batch electrocoagulation reactor. J. Hazard. Mater. 169, 1127–1133.
834 https://doi.org/10.1016/j.jhazmat.2009.04.089
Pr

835 Asaithambi, P., Govindarajan, R., Yesuf, M.B., Selvakumar, P., Alemayehu, E., 2021.
836 Investigation of direct and alternating current–electrocoagulation process for the
al

837 treatment of distillery industrial effluent: Studies on operating parameters. J. Environ.


838 Chem. Eng. 9, 104811. https://doi.org/10.1016/j.jece.2020.104811
rn

839 Asaithambi, P., Sajjadi, B., Abdul Aziz, A.R., Wan Daud, W.M.A. bin, 2016. Performance
u

840 evaluation of hybrid electrocoagulation process parameters for the treatment of distillery
841 industrial effluent. Process Saf. Environ. Prot. 104, 406–412.
Jo

842 https://doi.org/10.1016/j.psep.2016.09.023
843 Ashraf, S.N., Rajapakse, J., Dawes, L.A., Millar, G.J., 2021. Impact of turbidity, hydraulic
844 retention time, and polarity reversal upon iron electrode based electrocoagulation pre-
845 treatment of coal seam gas associated water. Environ. Technol. Innov. 23, 101622.
846 https://doi.org/10.1016/j.eti.2021.101622
847 Ashraf, S.N., Rajapakse, J., Dawes, L.A., Millar, G.J., 2019. Electrocoagulation for the
848 purification of highly concentrated brine produced from reverse osmosis desalination of
849 coal seam gas associated water. J. Water Process Eng. 28, 300–310.
850 https://doi.org/10.1016/j.jwpe.2019.02.003
851 Bandaru, S.R.S., Roy, A., Gadgil, A.J., van Genuchten, C.M., 2020. Long-term electrode
852 behavior during treatment of arsenic contaminated groundwater by a pilot-scale iron
853 electrocoagulation system. Water Res. 175, 115668.
854 https://doi.org/10.1016/j.watres.2020.115668
855 Behbahani, M., Alavi Moghaddam, M.R., Arami, M., 2011a. Techno-economical evaluation
856 of fluoride removal by electrocoagulation process: Optimization through response

49
857 surface methodology. Desalination 271, 209–218.
858 https://doi.org/10.1016/j.desal.2010.12.033
859 Behbahani, M., Alavi Moghaddam, M.R., Mokhtar, A., 2011b. A Comparison Between
860 Aluminum And Iron Electrodes On Removal Of Phosphate From Aqueous Solutions By
861 Electrocoagulation Process. Int. J. Environ. Res. 5, 403–412.
862 Betancor-Abreu, A., Mena, V.F., González, S., Delgado, S., Souto, R.M., Santana, J.J., 2019.
863 Design and optimization of an electrocoagulation reactor for fluoride remediation in
864 underground water sources for human consumption. J. Water Process Eng. 31, 100865.
865 https://doi.org/10.1016/j.jwpe.2019.100865
866 Bian, Y., Ge, Z., Albano, C., Lobo, F.L., Ren, Z.J., 2019. Oily bilge water treatment using
867 DC/AC powered electrocoagulation. Environ. Sci. Water Res. Technol. 5, 1654–1660.
868 https://doi.org/10.1039/C9EW00497A

f
oo
869 Cerqueira, A.A., Souza, P.S.A., Marques, M.R.C., 2014. Effects of direct and alternating
870 current on the treatment of oily water in an electroflocculation process. Braz. J. Chem.
871 Eng. 31, 693–701. https://doi.org/10.1590/0104-6632.20140313s00002363
872
pr
Cesar Lopes Geraldino, H., Izabelle Simionato, J., Karoliny Formicoli de Souza Freitas, T.,
e-
873 Carla Garcia, J., Evelázio de Souza, N., 2016. Evaluation of the electrode wear and the
874 residual concentration of iron in a system of electrocoagulation. Desalination Water
Pr

875 Treat 57, 13377–13387. https://doi.org/10.1080/19443994.2015.1058192


876 Chen, X., Ren, P., Li, T., Trembly, J.P., Liu, X., 2018. Zinc removal from model wastewater
877 by electrocoagulation: Processing, kinetics and mechanism. Chem. Eng. J. 349, 358–
al

878 367. https://doi.org/10.1016/j.cej.2018.05.099


rn

879 Chen, Y.M., Zhou, B.X., Li, L.H., Song, Y.H., Li, J.H., Liu, Y.B., Cai, W.M., 2011.
880 Application of Pulse Electrocoagulation to Dye Wastewater Treatment. Adv. Mater.
u

881 Res. 233–235, 444–451. https://doi.org/10.4028/www.scientific.net/AMR.233-235.444


Jo

882 Chow, H., Ingelsson, M., Roberts, E.P.L., Pham, A.L.-T., 2021. How does periodic polarity
883 reversal affect the faradaic efficiency and electrode fouling during iron
884 electrocoagulation? Water Res. 203, 117497.
885 https://doi.org/10.1016/j.watres.2021.117497
886 Chow, H., Pham, A.L.-T., 2021. Mitigating Electrode Fouling in Electrocoagulation by
887 Means of Polarity Reversal: The Effects of Electrode Type, Current Density, and
888 Polarity Reversal Frequency. Water Res. 197, 117074.
889 https://doi.org/10.1016/j.watres.2021.117074
890 Dong, W., Gu, X., Shu, Y., Cao, D., Yu, J., Abdel-Fatah, M.A., Fu, H., 2022. Pulse
891 electrocoagulation combined with a coagulant to remove antimony in wastewater. J.
892 Water Process Eng. 47, 102749. https://doi.org/10.1016/j.jwpe.2022.102749
893 Donneys-Victoria, D., Bermúdez-Rubio, D., Torralba-Ramírez, B., Marriaga-Cabrales, N.,
894 Machuca-Martínez, F., 2019. Removal of indigo carmine dye by electrocoagulation
895 using magnesium anodes with polarity change. Environ. Sci. Pollut. Res. 26, 7164–
896 7176. https://doi.org/10.1007/s11356-019-04160-y

50
897 Donneys-Victoria, D., Marriaga-Cabrales, N., Machuca-Martínez, F., Benavides-Guerrero, J.,
898 Cloutier, S.G., 2020. Indigo carmine and chloride ions removal by electrocoagulation.
899 Simultaneous production of brucite and layered double hydroxides. J. Water Process
900 Eng. 33, 101106. https://doi.org/10.1016/j.jwpe.2019.101106
901 Dubrawski, K.L., van Genuchten, C.M., Delaire, C., Amrose, S.E., Gadgil, A.J., Mohseni,
902 M., 2015. Production and Transformation of Mixed-Valent Nanoparticles Generated by
903 Fe(0) Electrocoagulation. Environ. Sci. Technol. 49, 2171–2179.
904 https://doi.org/10.1021/es505059d
905 Dura, A., Breslin, C.B., 2019. The removal of phosphates using electrocoagulation with
906 Al−Mg anodes. J. Electroanal. Chem. 846, 113161.
907 https://doi.org/10.1016/j.jelechem.2019.05.043
908 Dutta, N., Gupta, A., 2022. Development of arsenic removal unit with electrocoagulation and

f
909 activated alumina sorption: Field trial at rural West Bengal, India. J. Water Process Eng.

oo
910 49, 103013. https://doi.org/10.1016/j.jwpe.2022.103013
911 Dutta, N., Haldar, A., Gupta, A., 2021. Electrocoagulation for Arsenic Removal: Field Trials
912
913 https://doi.org/10.1007/s00244-020-00799-8 pr
in Rural West Bengal. Arch. Environ. Contam. Toxicol. 80, 248–258.
e-
914 Ensano, B.M.B., Borea, L., Naddeo, V., Belgiorno, V., de Luna, M.D.G., Balakrishnan, M.,
Pr

915 Ballesteros, F.C., 2019. Applicability of the electrocoagulation process in treating real
916 municipal wastewater containing pharmaceutical active compounds. J. Hazard. Mater.
917 361, 367–373. https://doi.org/10.1016/j.jhazmat.2018.07.093
al

918 Eyvaz, M., 2016. Treatment of Brewery Wastewater with Electrocoagulation: Improving the
rn

919 Process Performance by Using Alternating Pulse Current. Int. J. Electrochem. Sci.
920 4988–5008. https://doi.org/10.20964/2016.06.11
u

921 Eyvaz, M., Kirlaroglu, M., Aktas, T.S., Yuksel, E., 2009. The effects of alternating current
Jo

922 electrocoagulation on dye removal from aqueous solutions. Chem. Eng. J. 153, 16–22.
923 https://doi.org/10.1016/j.cej.2009.05.028
924 Fekete, É., Lengyel, B., Cserfalvi, T., Pajkossy, T., 2016. Electrochemical dissolution of
925 aluminium in electrocoagulation experiments. J. Solid State Electrochem. 20, 3107–
926 3114. https://doi.org/10.1007/s10008-016-3195-6
927 Fu, S., Jia, H., Meng, X., Guo, Z., Wang, J., 2021. Fe-C micro-electrolysis-electrocoagulation
928 based on BFDA in the pre-treatment of landfill leachate: Enhanced mechanism and
929 electrode decay monitoring. Sci. Total Environ. 781, 146797.
930 https://doi.org/10.1016/j.scitotenv.2021.146797
931 Fu, Z., 2018. Treatment of Reactive Brilliant Blue X-BR Synthetic Wastewater by
932 Asynchronous Periodic Reversal Electrocoagulation and Its Strengthening Mechanism.
933 Water Air Soil Pollut. 229, 69. https://doi.org/10.1007/s11270-018-3707-3
934 Fuladpanjeh‐Hojaghan, B., Elsutohy, M.M., Kabanov, V., Heyne, B., Trifkovic, M., Roberts,
935 E.P.L., 2019. In‐Operando Mapping of pH Distribution in Electrochemical Processes.
936 Angew. Chem. Int. Ed. 58, 16815–16819. https://doi.org/10.1002/anie.201909238

51
937 Garcia-Segura, S., Eiband, M.M.S.G., de Melo, J.V., Martínez-Huitle, C.A., 2017.
938 Electrocoagulation and advanced electrocoagulation processes: A general review about
939 the fundamentals, emerging applications and its association with other technologies. J.
940 Electroanal. Chem. 801, 267–299. https://doi.org/10.1016/j.jelechem.2017.07.047
941 Gelover-Santiago, S.L., Pérez-Castrejón, S., Martín-Domínguez, A., Villegas-Mendoza, I.E.,
942 2012. Electrogeneration of aluminium to remove silica in water. Water Sci. Technol. 65,
943 434–439. https://doi.org/10.2166/wst.2012.865
944 Ghanizadeh, G., Shariati neghab, G., Salem, M., Khalagi, K., 2016. Taguchi experimental
945 design for electrocoagulation process using alternating and direct current on fluoride
946 removal from water. Desalin. Water Treat. 57, 12675–12683.
947 https://doi.org/10.1080/19443994.2015.1049562
948 Gobbi, L.C.A., Nascimento, I.L., Muniz, E.P., Rocha, S.M.S., Porto, P.S.S., 2018.

f
949 Electrocoagulation with polarity switch for fast oil removal from oil in water emulsions.

oo
950 J. Environ. Manage. 213, 119–125. https://doi.org/10.1016/j.jenvman.2018.01.069
951 Guo, Z., Zhang, Y., Jia, H., Guo, J., Meng, X., Wang, J., 2022. Electrochemical methods for
952
953 pr
landfill leachate treatment: A review on electrocoagulation and electrooxidation. Sci.
Total Environ. 806, 150529. https://doi.org/10.1016/j.scitotenv.2021.150529
e-
954 H. Nguyen, Q., 2020. Electrocoagulation with a nanosecond pulse power supply to remove
Pr

955 COD from municipal wastewater using iron electrodes. Int. J. Electrochem. Sci. 493–
956 504. https://doi.org/10.20964/2020.01.66
957 Haldar, A., Gupta, A., 2020. Application of electrocoagulation: issues with community-level
al

958 defluoridation. Int. J. Environ. Sci. Technol. 17, 789–798.


rn

959 https://doi.org/10.1007/s13762-019-02323-5
960 Hasani, G., Maleki, A., Daraei, H., Ghanbari, R., Safari, M., McKay, G., Yetilmezsoy, K.,
u

961 Ilhan, F., Marzban, N., 2019. A comparative optimization and performance analysis of
Jo

962 four different electrocoagulation-flotation processes for humic acid removal from
963 aqueous solutions. Process Saf. Environ. Prot. 121, 103–117.
964 https://doi.org/10.1016/j.psep.2018.10.025
965 He, C.-C., Hu, C.-Y., Lo, S.-L., 2016. Evaluation of sono-electrocoagulation for the removal
966 of Reactive Blue 19 passive film removed by ultrasound. Sep. Purif. Technol. 165, 107–
967 113. https://doi.org/10.1016/j.seppur.2016.03.047
968 Ibrahim, M.H., Moussa, D.T., El-Naas, M.H., Nasser, M.S., 2020. A perforated electrode
969 design for passivation reduction during the electrochemical treatment of produced water.
970 J. Water Process Eng. 33, 101091. https://doi.org/10.1016/j.jwpe.2019.101091
971 Ingelsson, M., Yasri, N., Roberts, E.P.L., 2020. Electrode passivation, faradaic efficiency,
972 and performance enhancement strategies in electrocoagulation—a review. Water Res.
973 187, 116433. https://doi.org/10.1016/j.watres.2020.116433
974 Jiang, B., Xin, S., Liu, Y., Nin, C., Bi, X., Xue, J., 2018. Energy-Efficient Electrochemical
975 Strategy for the Oxidative Sequestration of As(III) in Synthesized Anoxic Groundwater.
976 Ind. Eng. Chem. Res 57, 8068–8077. https://doi.org/10.1021/acs.iecr.8b01013

52
977 Kamaraj, R., Ganesan, P., Lakshmi, J., Vasudevan, S., 2013. Removal of copper from water
978 by electrocoagulation process—effect of alternating current (AC) and direct current
979 (DC). Environ. Sci. Pollut. Res. 20, 399–412. https://doi.org/10.1007/s11356-012-0855-
980 7
981 Karamati Niaragh, E., Alavi Moghaddam, M.R., Emamjomeh, M.M., 2017. Techno-
982 economical evaluation of nitrate removal using continuous flow electro-coagulation
983 process: optimization by Taguchi model. Water Sci. Technol.: Water Supply 17, 1703–
984 1711. https://doi.org/10.2166/ws.2017.073
985 Karamati-Niaragh, E., Alavi Moghaddam, M.R., Emamjomeh, M.M., Nazlabadi, E., 2019.
986 Evaluation of direct and alternating current on nitrate removal using a continuous
987 electrocoagulation process: Economical and environmental approaches through RSM. J.
988 Environ. Manag. 230, 245–254. https://doi.org/10.1016/j.jenvman.2018.09.091

f
989 Karimifard, S., Alavi Moghaddam, M.R., 2018. Application of response surface methodology

oo
990 in physicochemical removal of dyes from wastewater: A critical review. Sci. Total
991 Environ. 640–641, 772–797. https://doi.org/10.1016/j.scitotenv.2018.05.355
992
993 pr
Keshmirizadeh, E., Yousefi, S., Rofouei, M.K., 2011. An investigation on the new
operational parameter effective in Cr(VI) removal efficiency: A study on
e-
994 electrocoagulation by alternating pulse current. J. Hazard. Mater. 190, 119–124.
995 https://doi.org/10.1016/j.jhazmat.2011.03.010
Pr

996 Khaghani, R., Yousefi, N., Asgari, A.R., Kholdi Haghighi, R., Ghadiri, K., Arabloo, J., Ali,
997 F., Bagheri, A., Khazaei, M., Talebi, S.S., Khalili, F., Porfadakari, S., 2020. Malathion
al

998 removal by electrocoagulation process: iron and stainless-steel electrodes, direct and
999 alternating current and determining energy and electrode consumption and kinetic study.
rn

1000 Desalin. Water Treat. 201, 110–120. https://doi.org/10.5004/dwt.2020.25769


u

1001 Khandegar, V., Saroha, A.K., 2013. Electrocoagulation for the treatment of textile industry
1002 effluent – A review. J. Environ. Manag. 128, 949–963.
Jo

1003 https://doi.org/10.1016/j.jenvman.2013.06.043
1004 Kong, X., Zhou, Y., Xu, T., Hu, B., Lei, X., Chen, H., Yu, G., 2020. A novel technique of
1005 COD removal from electroplating wastewater by Fenton—alternating current
1006 electrocoagulation. Environ. Sci. Pollut. Res. 27, 15198–15210.
1007 https://doi.org/10.1007/s11356-020-07804-6
1008 Lobo, F.L., Wang, H., Huggins, T., Rosenblum, J., Linden, K.G., Ren, Z.J., 2016. Low-
1009 energy hydraulic fracturing wastewater treatment via AC powered electrocoagulation
1010 with biochar. J. Hazard. Mater. 309, 180–184.
1011 https://doi.org/10.1016/j.jhazmat.2016.02.020
1012 Madhavan, M.A., Antony, S.P., 2021. Effect of polarity shift on the performance of
1013 electrocoagulation process for the treatment of produced water. Chemosphere 263,
1014 128052. https://doi.org/10.1016/j.chemosphere.2020.128052
1015 Maher, E.K., O’Malley, K.N., Heffron, J., Huo, J., Mayer, B.K., Wang, Y., McNamara, P.J.,
1016 2019. Analysis of operational parameters, reactor kinetics, and floc characterization for

53
1017 the removal of estrogens via electrocoagulation. Chemosphere 220, 1141–1149.
1018 https://doi.org/10.1016/j.chemosphere.2018.12.161
1019 Mansoorian, H.J., Mahvi, A.H., Jafari, A.J., 2014. Removal of lead and zinc from battery
1020 industry wastewater using electrocoagulation process: Influence of direct and alternating
1021 current by using iron and stainless steel rod electrodes. Sep. Purif. Technol. 135, 165–
1022 175. https://doi.org/10.1016/j.seppur.2014.08.012
1023 Mansouri, K., Ibrik, K., Bensalah, N., Abdel-Wahab, A., 2011. Anodic Dissolution of Pure
1024 Aluminum during Electrocoagulation Process: Influence of Supporting Electrolyte,
1025 Initial pH, and Current Density. Ind. Eng. Chem. Res 50, 13362–13372.
1026 https://doi.org/10.1021/ie201206d
1027 Mao, X., Hong, S., Zhu, H., Lin, H., Wei, L., Gan, F., 2008. Alternating pulse current in
1028 electrocoagulation for wastewater treatment to prevent the passivation of al electrode. J.

f
1029 Wuhan Univ. Technol. Mater. Sci. Ed. 23, 239–241. https://doi.org/10.1007/s11595-

oo
1030 006-2239-7
1031 Mohora, E., Rončević, S., Agbaba, J., Tubić, A., Mitić, M., Klašnja, M., Dalmacija, B., 2014.
1032
1033 pr
Removal of arsenic from groundwater rich in natural organic matter (NOM) by
continuous electrocoagulation/flocculation (ECF). Sep. Purif. Technol. 136, 150–156.
e-
1034 https://doi.org/10.1016/j.seppur.2014.09.006
Pr

1035 Mollah, M., Morkovsky, P., Gomes, J., Kesmez, M., Parga, J., Cocke, D., 2004.
1036 Fundamentals, present and future perspectives of electrocoagulation. J. Hazard. Mater.
1037 114, 199–210. https://doi.org/10.1016/j.jhazmat.2004.08.009
al

1038 Moradi, M., Vasseghian, Y., Arabzade, H., Mousavi Khaneghah, A., 2021. Various
rn

1039 wastewaters treatment by sono-electrocoagulation process: A comprehensive review of


1040 operational parameters and future outlook. Chemosphere 263, 128314.
u

1041 https://doi.org/10.1016/j.chemosphere.2020.128314
Jo

1042 Müller, S., Behrends, T., van Genuchten, C.M., 2019. Sustaining efficient production of
1043 aqueous iron during repeated operation of Fe(0)-electrocoagulation. Water Res. 155,
1044 455–464. https://doi.org/10.1016/j.watres.2018.11.060
1045 Nazlabadi, E., Alavi Moghaddam, M.R., 2017. Simulation of Nitrate Removal in a Batch
1046 Flow Electrocoagulation-Flotation (ECF) Process by Response Surface Method (RSM),
1047 in: Trends in Asian Water Environmental Science and Technology. Springer
1048 International Publishing, Cham, pp. 49–60. https://doi.org/10.1007/978-3-319-39259-
1049 2_4
1050 Nazlabadi, E., Alavi Moghaddam, M.R., Karamati-Niaragh, E., 2019. Simultaneous removal
1051 of nitrate and nitrite using electrocoagulation/floatation (ECF): A new multi-response
1052 optimization approach. J. Environ. Manag. 250, 109489.
1053 https://doi.org/10.1016/j.jenvman.2019.109489
1054 Nguyen, Q.H., Watari, T., Yamaguchi, T., Kawamura, Y., Suematsu, H., Wiff, J.P., Niihara,
1055 K., Nakayama, T., 2020. Comparison between Nanosecond Pulse and Direct Current
1056 Electrocoagulation for Textile Wastewater Treatment. J. Water Environ. Technol. 18,
1057 147–156. https://doi.org/10.2965/jwet.19-080

54
1058 Nidheesh, P.V., Oladipo, A.A., Yasri, N.G., Laiju, A.R., Cheela, V.R.S., Thiam, A., Asfaha,
1059 Y.G., Kanmani, S., Roberts, E. (Ted) P.L., 2022. Emerging applications, reactor design
1060 and recent advances of electrocoagulation process. Process Saf. Environ. Prot. 166, 600–
1061 616. https://doi.org/10.1016/j.psep.2022.08.051
1062 Nidheesh, P.V., Scaria, J., Babu, D.S., Kumar, M.S., 2021. An overview on combined
1063 electrocoagulation-degradation processes for the effective treatment of water and
1064 wastewater. Chemosphere 263, 127907.
1065 https://doi.org/10.1016/j.chemosphere.2020.127907
1066 Oliveira, J.T., de Sousa, M.C., Martins, I.A., de Sena, L.M.G., Nogueira, T.R., Vidal, C.B.,
1067 Neto, E.F.A., Romero, F.B., Campos, O.S., do Nascimento, R.F., 2021.
1068 Electrocoagulation/oxidation/flotation by direct pulsed current applied to the removal of
1069 antibiotics from Brazilian WWTP effluents. Electrochim. Acta 388, 138499.
1070 https://doi.org/10.1016/j.electacta.2021.138499

f
oo
1071 Othmani, A., Kesraoui, A., Seffen, M., 2017. The alternating and direct current effect on the
1072 elimination of cationic and anionic dye from aqueous solutions by electrocoagulation

pr
1073 and coagulation flocculation. Euro-Mediterr. J. Environ. Integr. 2, 6.
1074 https://doi.org/10.1007/s41207-017-0016-y
e-
1075 Ozyonar, F., Karagozoglu, B., 2015. Treatment of pretreated coke wastewater by
1076 electrocoagulation and electrochemical peroxidation processes. Sep. Purif. Technol. 150,
Pr

1077 268–277. https://doi.org/10.1016/j.seppur.2015.07.011


1078 Payami Shabestar, M., Alavi Moghaddam, M.R., Karamati-Niaragh, E., 2021. Evaluation of
al

1079 energy and electrode consumption of Acid Red 18 removal using electrocoagulation
1080 process through RSM: alternating and direct current. Environ. Sci. Pollut. Res. 28,
rn

1081 67214–67223. https://doi.org/10.1007/s11356-021-15345-9


u

1082 Pertile, T.S., Birriel, E.J., 2017. Treatment of hydrocyanic galvanic effluent by
1083 electrocoagulation: Optimization of operating parameters using statistical techniques and
Jo

1084 a coupled polarity inverter. Korean J. Chem. Eng. 34, 2631–2640.


1085 https://doi.org/10.1007/s11814-017-0178-y
1086 Pi, K.-W., Xiao, Q., Zhang, H.-Q., Xia, M., Gerson, A.R., 2014. Decolorization of synthetic
1087 Methyl Orange wastewater by electrocoagulation with periodic reversal of electrodes
1088 and optimization by RSM. Process Saf. Environ. Prot. 92, 796–806.
1089 https://doi.org/10.1016/j.psep.2014.02.008
1090 Punckt, C., Bölscher, M., Rotermund, H.H., Mikhailov, A.S., Organ, L., Budiansky, N.,
1091 Scully, J.R., Hudson, J.L., 2004. Sudden Onset of Pitting Corrosion on Stainless Steel as
1092 a Critical Phenomenon. Science (1979) 305, 1133–1136.
1093 https://doi.org/10.1126/science.1101358
1094 Que, N.H., Kawamura, Y., Watari, T., Takimoto, Y., Yamaguchi, T., Suematsu, H., Niihara,
1095 K., Wiff, J.P., Nakayama, T., 2021. Nanosecond pulse used to enhance the
1096 electrocoagulation of municipal wastewater treatment with low specific energy
1097 consumption. Environ. Technol. 42, 2154–2162.
1098 https://doi.org/10.1080/09593330.2019.1694082

55
1099 Rajaei, F., Taheri, E., Hadi, S., Fatehizadeh, A., Amin, M.M., Rafei, N., Fadaei, S.,
1100 Aminabhavi, T.M., 2021. Enhanced removal of humic acid from aqueous solution by
1101 combined alternating current electrocoagulation and sulfate radical. Environ. Pollut.
1102 277, 116632. https://doi.org/10.1016/j.envpol.2021.116632
1103 Ren, M., Song, Y., Xiao, S., Zeng, P., Peng, J., 2011. Treatment of berberine hydrochloride
1104 wastewater by using pulse electro-coagulation process with Fe electrode. Chem. Eng. J.
1105 169, 84–90. https://doi.org/10.1016/j.cej.2011.02.056
1106 Roberge, P.R., 2008. Corrosion Engineering: Principles and Practice. McGraw-Hill
1107 Education.
1108 Rodrigues, C., Follmann, H.V.D.M., Núñez-Gómez, D., Nagel-Hassemer, M.E., Lapolli,
1109 F.R., Lobo-Recio, M.Á., 2020. Sulfate removal from mine-impacted water by
1110 electrocoagulation: statistical study, factorial design, and kinetics. Environ. Sci. Pollut.

f
1111 Res. 27, 39572–39583. https://doi.org/10.1007/s11356-020-09758-1

oo
1112 Rodrigues Pires da Silva, J., Merçon, F., Firmino da Silva, L., Andrade Cerqueira, A., Braz
1113 Ximango, P., da Costa Marques, M.R., 2015. Evaluation of electrocoagulation as pre-
1114
1115 135. https://doi.org/10.1016/j.jwpe.2015.09.009 pr
treatment of oil emulsions, followed by reverse osmosis. J. Water Process Eng. 8, 126–
e-
1116 Sadeghi, H., Mohammadpour, A., Samaei, M.R., Azhdarpoor, A., Hadipoor, M., Mehrazmay,
Pr

1117 H., Mousavi Khaneghah, A., 2022. Application of sono-electrocoagulation in arsenic


1118 removal from aqueous solutions and the related human health risk assessment. Environ.
1119 Res. 212, 113147. https://doi.org/10.1016/j.envres.2022.113147
al

1120 Sanei, E., Mokhtarani, N., 2022. Leachate post-treatment by electrocoagulation process:
rn

1121 Effect of polarity switching and anode-to-cathode surface area. J. Environ. Manage. 319,
1122 115733. https://doi.org/10.1016/j.jenvman.2022.115733
u

1123 Secula, M., Cretescu, I., Cagnon, B., Manea, L., Stan, C., Breaban, I., 2013. Fractional
Jo

1124 Factorial Design Study on the Performance of GAC-Enhanced Electrocoagulation


1125 Process Involved in Color Removal from Dye Solutions. Materials 6, 2723–2746.
1126 https://doi.org/10.3390/ma6072723
1127 Sefatjoo, P., Alavi Moghaddam, M.R., Mehrabadi, A.R., 2020. Evaluating electrocoagulation
1128 pretreatment prior to reverse osmosis system for simultaneous scaling and colloidal
1129 fouling mitigation: Application of RSM in performance and cost optimization. J. Water
1130 Process Eng. 35, 101201. https://doi.org/10.1016/j.jwpe.2020.101201
1131 Shokri, A., Fard, M.S., 2022. A critical review in electrocoagulation technology applied for
1132 oil removal in industrial wastewater. Chemosphere 288, 132355.
1133 https://doi.org/10.1016/j.chemosphere.2021.132355
1134 Syam Babu, D., Vijay, K., Nidheesh, P.V., Suresh Kumar, M., 2021. Performance of
1135 continuous aerated iron electrocoagulation process for arsenite removal from simulated
1136 groundwater and management of arsenic-iron sludge. Sustain. Energy Technol. Assess.
1137 47, 101476. https://doi.org/10.1016/j.seta.2021.101476
1138 Taheri, M., Alavi Moghaddam, M.R., Arami, M., 2013. Techno-economical optimization of
1139 Reactive Blue 19 removal by combined electrocoagulation/coagulation process through

56
1140 MOPSO using RSM and ANFIS models. J. Environ. Manag. 128, 798–806.
1141 https://doi.org/10.1016/j.jenvman.2013.06.029
1142 Taheri, M., Alavi Moghaddam, M.R., Arami, M., 2012. Optimization of Acid Black 172
1143 decolorization by electrocoagulation using response surface methodology. J. Environ.
1144 Health Sci. Eng. 9, 23. https://doi.org/10.1186/1735-2746-9-23
1145 Tegladza, I.D., Xu, Q., Xu, K., Lv, G., Lu, J., 2021. Electrocoagulation processes: A general
1146 review about role of electro-generated flocs in pollutant removal. Process Saf. Environ.
1147 Prot. 146, 169–189. https://doi.org/10.1016/j.psep.2020.08.048
1148 Timmes, T.C., Kim, H.-C., Dempsey, B.A., 2010. Electrocoagulation pretreatment of
1149 seawater prior to ultrafiltration: Pilot-scale applications for military water purification
1150 systems. Desalination 250, 6–13. https://doi.org/10.1016/j.desal.2009.03.021
1151 Trompette, J.-L., Lahitte, J.-F., 2021. Effects of some ion-specific properties in the

f
oo
1152 electrocoagulation process with aluminum electrodes. Colloids Surf. A Physicochem.
1153 Eng. Asp. 629, 127507. https://doi.org/10.1016/j.colsurfa.2021.127507
1154
1155
pr
van Genuchten, C.M., Bandaru, S.R.S., Surorova, E., Amrose, S.E., Gadgil, A.J., Peña, J.,
2016. Formation of macroscopic surface layers on Fe(0) electrocoagulation electrodes
e-
1156 during an extended field trial of arsenic treatment. Chemosphere 153, 270–279.
1157 https://doi.org/10.1016/j.chemosphere.2016.03.027
Pr

1158 van Genuchten, C.M., Dalby, K.N., Ceccato, M., Stipp, S.L.S., Dideriksen, K., 2017. Factors
1159 affecting the Faradaic efficiency of Fe(0) electrocoagulation. J. Environ. Chem. Eng. 5,
1160 4958–4968. https://doi.org/10.1016/j.jece.2017.09.008
al

1161 Vasudevan, S., 2012. Effects of alternating current (AC) and direct current (DC) in
rn

1162 electrocoagulation process for the removal of iron from water. Can. J. Chem. Eng. 90,
1163 1160–1169. https://doi.org/10.1002/cjce.20625
u

Vasudevan, S., Kannan, B.S., Lakshmi, J., Mohanraj, S., Sozhan, G., 2011a. Effects of
Jo

1164
1165 alternating and direct current in electrocoagulation process on the removal of fluoride
1166 from water. J. Chem. Technol. Biotechnol. 86, 428–436.
1167 https://doi.org/10.1002/jctb.2534
1168 Vasudevan, S., Lakshmi, J., 2012. Effect of alternating and direct current in an
1169 electrocoagulation process on the removal of cadmium from water. Water Sci. Technol.
1170 65, 353–360. https://doi.org/10.2166/wst.2012.859
1171 Vasudevan, S., Lakshmi, J., 2011. Effects of alternating and direct current in
1172 electrocoagulation process on the removal of cadmium from water – A novel approach.
1173 Sep. Purif. Technol. 80, 643–651. https://doi.org/10.1016/j.seppur.2011.06.027
1174 Vasudevan, S., Lakshmi, J., Sozhan, G., 2012. Studies on the removal of arsenate from water
1175 through electrocoagulation using direct and alternating current. Desalin. Water Treat. 48,
1176 163–173. https://doi.org/10.1080/19443994.2012.698809
1177 Vasudevan, S., Lakshmi, J., Sozhan, G., 2011b. Effects of alternating and direct current in
1178 electrocoagulation process on the removal of cadmium from water. J. Hazard. Mater.
1179 192, 26–34. https://doi.org/10.1016/j.jhazmat.2011.04.081

57
1180 Wang, T., Liu, T., 2017. Pulse electro-coagulation application in treating dibutyl phthalate
1181 wastewater. Water Sci. Technol. 76, 1124–1131.
1182 Wellner, D.B., Couperthwaite, S.J., Millar, G.J., 2018. Influence of operating parameters
1183 during electrocoagulation of sodium chloride and sodium bicarbonate solutions using
1184 aluminium electrodes. J. Water Process Eng. 22, 13–26.
1185 https://doi.org/10.1016/j.jwpe.2017.12.014
1186 Xin, S., Nin, C., Gong, Y., Ma, J., Bi, X., Jiang, B., 2018. A full-wave rectified alternating
1187 current wireless electrocoagulation strategy for the oxidative remediation of As(III) in
1188 simulated anoxic groundwater. Chem. Eng. J. 351, 1047–1055.
1189 https://doi.org/10.1016/j.cej.2018.06.178
1190 Xu, H., Yang, Z., Luo, Y., Zeng, G., Huang, J., Wang, L., Song, P., Yang, X., 2015. A novel
1191 approach to sustain Fe 0 -electrocoagulation for Cr(VI) removal by optimizing chloride

f
1192 ions. Sep. Purif. Technol. 156, 200–206. https://doi.org/10.1016/j.seppur.2015.09.074

oo
1193 Xu, L., Huang, Q., Xu, X., Cao, G., He, C., Wang, Y., Yang, M., 2017. Simultaneous
1194 removal of Zn 2+ and Mn 2+ ions from synthetic and real smelting wastewater using
1195
1196 pr
electrocoagulation process: Influence of pulse current parameters and anions. Sep. Purif.
Technol. 188, 316–328. https://doi.org/10.1016/j.seppur.2017.07.036
e-
1197 Xu, L., Xu, X., Cao, G., Liu, S., Duan, Z., Song, S., Song, M., Zhang, M., 2018.
Pr

1198 Optimization and assessment of Fe–electrocoagulation for the removal of potentially


1199 toxic metals from real smelting wastewater. J. Environ. Manag. 218, 129–138.
1200 https://doi.org/10.1016/j.jenvman.2018.04.049
al

1201 Xu, T., Zheng, X., Zhou, Y., Zhu, C., Hu, B., Lei, X., Zhang, X., Yu, G., 2021. Study on the
rn

1202 treatment of Cu2+-organic compound wastewater by electro-Fenton coupled pulsed AC


1203 coagulation. Chemosphere 280, 130679.
u

1204 https://doi.org/10.1016/j.chemosphere.2021.130679
Jo

1205 Xu, T., Zhou, Y., Hu, B., Lei, X., Yu, G., 2020. Comparison between sinusoidal AC
1206 coagulation and conventional DC coagulation in removing Cu2+ from printed circuit
1207 board wastewater. Ecotoxicol. Environ. Saf. 197, 110629.
1208 https://doi.org/10.1016/j.ecoenv.2020.110629
1209 Xu, T., Zhou, Y., Lei, X., Hu, B., Chen, H., Yu, G., 2019. Study on highly efficient Cr(VI)
1210 removal from wastewater by sinusoidal alternating current coagulation. J. Environ.
1211 Manag. 249, 109322. https://doi.org/10.1016/j.jenvman.2019.109322
1212 Yaghmaeian, K., Mahvi, A., Nasseri, S., Hooshangi Shayesteh, M., Jafari Mansoorian, H.,
1213 Khanjani, N., 2020. Drinking water softening with electrocoagulation process: Influence
1214 of direct and alternating currents as inductive with different arrangement rod electrodes
1215 and polarity inverter. Sci. Iran. 27, 1275–1292.
1216 Yang, J., Liu, F., Bu, Y., Wei, N., Liu, S., Chang, J., Chen, X., Zhang, W., Zhou, R., Zhang,
1217 C., 2020. Cd removal by direct and positive single pulse current electrocoagulation:
1218 Operating conditions and energy consumption. Environ. Technol. Innov. 20, 101123.
1219 https://doi.org/10.1016/j.eti.2020.101123

58
1220 Yang, Z., Xu, H., Zeng, G., Luo, Y., Yang, X., Huang, J., Wang, L., Song, P., 2015. The
1221 behavior of dissolution/passivation and the transformation of passive films during
1222 electrocoagulation: Influences of initial pH, Cr(VI) concentration, and alternating pulsed
1223 current. Electrochim. Acta 153, 149–158. https://doi.org/10.1016/j.electacta.2014.11.183
1224 Yasasve, M., Manjusha, M., Manojj, D., Hariharan, N.M., Sai Preethi, P., Asaithambi, P.,
1225 Karmegam, N., Saravanan, M., 2022. Unravelling the emerging carcinogenic
1226 contaminants from industrial waste water for prospective remediation by
1227 electrocoagulation – A review. Chemosphere 307, 136017.
1228 https://doi.org/10.1016/j.chemosphere.2022.136017
1229 Yasri, N.G., Ingelsson, M., Nightingale, M., Jaggi, A., Dejak, M., Kryst, K., Oldenburg,
1230 T.B.P., Roberts, E.P.L., 2022. Investigation of electrode passivation during
1231 electrocoagulation treatment with aluminum electrodes for high silica content produced
1232 water. Water Sci. Technol. 85, 925–942. https://doi.org/10.2166/wst.2022.012

f
oo
1233 Yu, Y., Zhong, Y., Wang, M., Guo, Z., 2021. Electrochemical behavior of aluminium anode
1234 in super-gravity field and its application in copper removal from wastewater by

pr
1235 electrocoagulation. Chemosphere 272, 129614.
1236 https://doi.org/10.1016/j.chemosphere.2021.129614
e-
1237 Zaleschi, L., Secula, M.S., Teodosiu, C., Stan, C.S., Cretescu, I., 2014. Removal of
1238 Rhodamine 6G from Aqueous Effluents by Electrocoagulation in a Batch Reactor:
Pr

1239 Assessment of Operational Parameters and Process Mechanism. Water Air Soil Pollut.
1240 225, 2101. https://doi.org/10.1007/s11270-014-2101-z
al

1241 Zhang, B., Ma, X.L., 2019. A review—Pitting corrosion initiation investigated by TEM. J.
1242 Mater. Sci. Technol. 35, 1455–1465. https://doi.org/10.1016/j.jmst.2019.01.013
rn

1243 Zhang, H., Bian, J., Yang, C., Hu, Z., Liu, F., Zhang, C., 2022. Removal of tetracycline from
u

1244 livestock wastewater by positive single pulse current electrocoagulation: Mechanism,


1245 toxicity assessment and cost evaluation. Sci. Total Environ. 810, 151955.
Jo

1246 https://doi.org/10.1016/j.scitotenv.2021.151955
1247 Zhang, J., Li, J., Ma, C., Yi, L., Gu, T., Wang, J., 2021. High-efficiency and energy-saving
1248 alternating pulse current electrocoagulation to remove polyvinyl alcohol in wastewater.
1249 RSC Adv. 11, 40085–40099. https://doi.org/10.1039/D1RA08093H
1250 Zhang, W., Zhou, Y., Hu, C., Qu, J., 2021. Electricity generation from salinity gradient to
1251 remove chromium using reverse electrodialysis coupled with electrocoagulation.
1252 Electrochim. Acta 379, 138153. https://doi.org/10.1016/j.electacta.2021.138153
1253 Zheng, T., 2017. Treatment of oilfield produced water with electrocoagulation: improving the
1254 process performance by using pulse current. J. Water Reuse Desalin. 7, 378–386.
1255 https://doi.org/10.2166/wrd.2016.113
1256 Zhou, L., Liu, D., Li, S., Yin, X., Zhang, Chunlei, Li, X., Zhang, Chuguo, Zhang, W., Cao,
1257 X., Wang, J., Wang, Z.L., 2019. Effective removing of hexavalent chromium from
1258 wasted water by triboelectric nanogenerator driven self-powered electrochemical system
1259 – Why pulsed DC is better than continuous DC? Nano Energy 64, 103915.
1260 https://doi.org/10.1016/j.nanoen.2019.103915

59
1261 Zhou, R., Liu, F., Wei, N., Yang, C., Yang, J., Wu, Y., Li, Y., Xu, K., Chen, X., Zhang, C.,
1262 2020. Comparison of Cr(VI) removal by direct and pulse current electrocoagulation:
1263 Implications for energy consumption optimization, sludge reduction and floc
1264 magnetism. J. Water Process Eng. 37, 101387.
1265 https://doi.org/10.1016/j.jwpe.2020.101387
1266 Zhou, Y., Chen, S., Qiu, J., Zhu, C., Xu, T., Zeng, M., He, X., Hu, B., Zhang, X., Yu, G.,
1267 2021. Removal of phosphorus in wastewater by sinusoidal alternating current
1268 coagulation: performance and mechanism. Environ. Technol. 1–14.
1269 https://doi.org/10.1080/09593330.2021.1916093
1270 Zhu, H., Wang, Q., Zhang, F., Yang, C., Li, Y., Zhou, C., 2022. Fuzzy comprehensive
1271 evaluation strategy for operating state of electrocoagulation purification process based
1272 on sliding window. Process Saf. Environ. Prot 165, 217–229.
1273 https://doi.org/10.1016/j.psep.2022.06.063

f
oo
1274

pr
e-
Pr
al
u rn
Jo

60
Highlights

▪ Electrode passivation as a detrimental phenomenon to the EC process is outlined

▪ Mechanisms involved in improved EC under various current waveforms are discussed

▪ The PDC, APC, and AC waveforms as alternatives to conventional DC are summarized

▪ The role of these waveforms in de-passivation and enhancing performance is explored

▪ How current waveform parameters affect PDC-EC and APC-EC performance is

discussed

f
oo
pr
e-
Pr
al
u rn
Jo
Declaration of interests

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

☐ The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

You might also like