You are on page 1of 33

Critical Reviews in Environmental Science and

Technology

ISSN: 1064-3389 (Print) 1547-6537 (Online) Journal homepage: http://www.tandfonline.com/loi/best20

Electro-Fenton process for water and wastewater


treatment

Huanqi He & Zhi Zhou

To cite this article: Huanqi He & Zhi Zhou (2017): Electro-Fenton process for water and
wastewater treatment, Critical Reviews in Environmental Science and Technology, DOI:
10.1080/10643389.2017.1405673

To link to this article: https://doi.org/10.1080/10643389.2017.1405673

Published online: 18 Dec 2017.

Submit your article to this journal

View related articles

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


http://www.tandfonline.com/action/journalInformation?journalCode=best20

Download by: [University of Florida] Date: 19 December 2017, At: 00:51


CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY
2017, VOL. 0, NO. 0, 1–32
https://doi.org/10.1080/10643389.2017.1405673

Electro-Fenton process for water and wastewater treatment


a,b
Huanqi Hea and Zhi Zhou
a
Lyles School of Civil Engineering, Purdue University, West Lafayette, Indiana, USA; bDivision of
Environmental and Ecological Engineering, Purdue University, West Lafayette, Indiana, USA

ABSTRACT KEYWORDS
Electro-Fenton process is an emerging treatment technology for electro-Fenton; hydroxyl
water and wastewater treatment. Electro-Fenton utilizes hydroxyl radical; water treatment;
wastewater treatment
Downloaded by [University of Florida] at 00:51 19 December 2017

radicals to oxidize hazardous contaminants and is especially


useful to treat recalcitrant compounds that are not easily
degraded in conventional water and wastewater treatment plants.
This paper reviews the mechanism of electro-Fenton process and
its kinetic modeling. The effects of electrode materials, hydrogen
peroxide concentration, ferrous ion concentration, pH, electrolyte,
oxygen sparging rate, temperature, initial concentration of
pollutant, feeding mode, current density, and distance between
electrodes are also reviewed. The performance of electro-Fenton
is compared with other treatment technologies. Finally, existing
challenges and future perspectives are discussed.

1. Introduction
The wide occurrence of organic contaminants in agricultural runoff, domestic sew-
age, industrial wastewater, and polluted soils poses a severe threat to public health
as well as our ecosystem. Many of those contaminants are extremely toxic even at
low concentrations, and therefore need to be effectively removed. However, many
refractory organic contaminants, especially aromatic compounds, can hardly be
removed by commonly employed treatment technologies in conventional biologi-
cal wastewater treatment plants (Annabi et al., 2016; Lin & Lo, 1997). Therefore,
an efficient treatment technology for recalcitrant contaminants is needed.
Advanced oxidation processes (AOPs) are powerful treatment technologies
for effective removal of refractory organic contaminants through oxidation with
hydroxyl radicals (OH). Development of AOPs during the last few decades
allows scientists and engineers to gain an in-depth understanding of their oxi-
dation mechanisms and to enhance their performance for contaminant
removal. Among AOPs, Fenton technology is very attractive due to its simplic-
ity, low cost, high performance, and the lack of toxicity of the Fenton’s reagents

CONTACT Zhi Zhou zhizhou@purdue.edu Division of Environmental and Ecological Engineering, Purdue
University, West Lafayette, IN 47907, USA.
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/best.
© 2017 Taylor & Francis Group, LLC
2 H. HE AND Z. ZHOU

(ferrous ion and hydrogen peroxide) (Nidheesh & Gandhimathi, 2012; Wang,
Zheng, Zhang, & Wang, 2016). During the Fenton process, ferrous ion (Fe2C)
catalyzes hydrogen peroxide (H2O2) into OH, which is a highly reactive and
strong oxidizing agent that can react with most organic compounds containing
C-H and C-C bonds at near diffusion-controlled rates (>10¡9 M¡1S¡1)
(B€uy€uks€onmez, Rynk, Hess, & Bechinski, 1999; Haag & Yao, 1992). Fenton’s
reagents remove organic contaminants under normal temperature and pressure
conditions and allows high depuration with relatively inexpensive and abundant
materials (Duesterberg & Waite, 2006). In addition, both ferrous and ferric
(Fe3C) ions are coagulants, and coagulation during Fenton processes can further
remove organic contaminants (Badawy & Ali, 2006). With the extraordinary
oxidation performance and fundamental understanding of the chemical mecha-
nism, Fenton process has been considered as one of the most attractive AOPs
Downloaded by [University of Florida] at 00:51 19 December 2017

and has been applied in water and wastewater treatment, biomedical systems,
atmospheric processes, and biogeochemistry (Faust & Hoigne, 1990; Zuo &
Hoigne, 1992).
However, there are still a few challenges in conventional Fenton-based sys-
tems, including storage and transition of highly concentrated H2O2, rapid con-
sumption of catalysts, production and additional disposal of generated iron
sludge (Ma, Zhou, Ren, Yang, & Liang, 2016; Zhang, Zhang, & Zhou, 2006).
Therefore, efforts have been made to develop new technologies to address these
challenges while still utilizing the strong oxidation efficiency of Fenton process.
Electro-Fenton process has been a new development of Fenton process and has
raised a great interest for the removal of organic contaminants. In an electro-
Fenton process, contaminants are removed by Fenton’s reagents together with
anodic oxidation on the anode surface. Due to the formation of refractory car-
boxylic acids, anodic oxidation alone is not effective to mineralize most aromatic
pollutants (Babuponnusami & Muthukumar, 2014). But with the generation of
OH, electro-Fenton process can reach a notable mineralization efficiency of
organic contaminants. Compared with the conventional Fenton process, electro-
Fenton process avoids the transport and storage of external H2O2 with in situ
generated H2O2 (Gao, Zhang, Hao, & Vecitis, 2015; Liu, Xie, Ong, Vecitis, &
Zhou, 2015; Rosales, Pazos, & Sanroman, 2012), and therefore is an environmen-
tally friendly technology as chemical usage has been reduced.
Although electro-Fenton offers significant advantages over conventional Fenton
process, slow production of H2O2 due to the low solubility of oxygen, low current
efficiency at higher pH, and higher operational costs have limited the application
of electro-Fenton. To fully understand the mechanism of electro-Fenton process, a
comprehensive literature review is needed to help us better understand the assets
and challenges of this technology. This review covers the recent advancements in
the mechanism, modeling, impacting factors of electro-Fenton process and identi-
fies challenges and opportunities of electro-Fenton technique.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 3

2. Mechanism
The main mechanism of a Fenton system is based on the oxidation of OH, which
is one of the most powerful oxidizing agents. In a Fenton system, OH is produced
via a homogeneous Fenton reaction (Eq. 1) and serves as the initiating reaction of
the chain Fenton reactions (Eqs. 1–5) (Kremer, 1999).

H2 O2 C Fe2 C ! Fe3 C C OH C OH ¡ (1)


 
OH C H2 O2 ! HO2 C H2 O (2)
3C  2C C
Fe C HO2 ! Fe C H C O2 (3)
2C  3C ¡
Fe C HO2 ! Fe C HO2 (4)
2C  3C ¡
Fe C OH ! Fe C OH (5)
Downloaded by [University of Florida] at 00:51 19 December 2017

According to the free radical theory (Barb, Baxendale, George, & Hargrave,
1951), Eqs. (1–3) forms the cycle where O2 is released and Eqs. (4) and (5)
act as termination reactions (Kremer, 1999). The active intermediate species,
OH, is recognized as a strong oxidizing agent that can efficiently degrade
most organic compounds into harmless compounds via dehydrogenation or
hydroxylation (Borras, Arias, Oliver, & Brillas, 2011; Ribeiro, Nunes, Pereira,
& Silva, 2015; Tokumura, Sugawara, Raknuzzaman, Habibullah-Al-Mamun, &
Masunaga, 2016).
Besides the basic Fenton reactions, the commonly accepted mechanism of an
electro-Fenton process includes in situ generation of H2O2 on the cathode
(Eq. 6) and sacrificial production of Fe2C on the anode (Eq. 7) (Qiu, He, Ma,
Liu, & Waite, 2015). No matter whether the Fenton’s reagents are externally
applied or in situ generated in a sacrificial anode, the homogeneous Fenton
reaction (Eq. 1) produces the active oxidizing agent OH. Interestingly, in elec-
tro-Fenton systems, OH generation was also observed at the anode by the oxi-
dation of water (Eq. 8) (Garcia, Isarain-Chavez, Garcia-Segura, Brillas, &
Peralta-Hernandez, 2013; Qiu et al., 2015), but the reaction of water discharge
is very slow (k D 10¡10 Ms¡1 at pH 3.0 (Qiu et al., 2015)), requiring high cur-
rents and suitable electrode materials.

Cathode : O2 C 2H C C 2e ¡ ! H2 O2 (6)
2C ¡
Sacrificial Anode : Fe ! Fe 0
Ce (7)
H2 O ! OH C H C C e ¡ (8)

OH is a second strongest oxidizing agent preceded by F2 (Table 1). OH has a
standard potential as high as 2.80 V (vs. SHE), indicating its extraordinary ability
to obtain electrons from other materials and to oxidize other substances.
Eqs. (9–13) illustrate the degradation mechanism where RH denotes organic
compounds. Hydroxyl radicals react rapidly with RH and start a radical
4 H. HE AND Z. ZHOU

Table 1. Standard potential of some oxidizing agents.


Oxidizing Agent Standard Potential (V vs. SHE) Reference

Oxygen (g) (O2) 1.226 Brezonik and Arnold (2011)


Oxygen (aq) (O2) 1.268 Brezonik and Arnold (2011)
Chlorine (aq) (Cl2) 1.392 Brezonik and Arnold (2011)
Hypochlorous acid (HClO) 1.481 Brezonik and Arnold (2011)
Permanganate (MnO4)¡ 1.508 Brezonik and Arnold (2011)
Hydrogen peroxide (aq) (H2O2) 1.758 Brezonik and Arnold (2011)
Ozone (O3) 2.08 Sires et al. (2014)
Oxygen (atomic)(O) 2.42 Sires et al. (2014)
Hydroxyl radical (OH) 2.80 Sires et al. (2014)
Fluorine (F2) 3.06 Sires et al. (2014)

oxidation chain mainly by abstracting a hydrogen atom from C-H, N-H, or O-


H bonds (Eqs. 9 and 11), or adding to an unsaturated bond such as C═C
Downloaded by [University of Florida] at 00:51 19 December 2017

bonds or aromatic rings (ArH) (Eqs. 12,13) (Pignatello, Oliveros, & Mackay,
2006, Sires, Brillas, Oturan, Rodrigo, & Panizza, 2014). The oxidation of
organic compounds with electro-Fenton can be shown in the following exam-
ples. For perfluorooctanoate (PFOA, C7F15COOH) (Figure 1), researchers pro-
posed that OH attacked C7F15 that was generated from C7F15COO¡ and
formed C7F15OH. The formation of C6F13COO¡ was resulted from C7F15OH
after intramolecular rearrangement and the produced shorter PFOA chain fol-
lowed the same reactions as those of C7F15COO¡ until complete mineralization
(Liu et al., 2015). For the example of acid red 97 (AR97, C32H20N4S2O8Na2)
(Figure 2) (Kayan, G€ ozmen, Demirel, & Gizir, 2010), OH was observed to
attack the azo bonds (¡N═N¡) initially, and then degraded the products
(I–IV) to hydroxylated or poly hydroxylated derivatives via hydroxylation.

Figure 1. Schematic pathway for PFOA mineralization by electro-Fenton (Liu et al., 2015). Copyright
© 2015 American Chemical Society.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 5
Downloaded by [University of Florida] at 00:51 19 December 2017

Figure 2. Schematic pathway for AR97 mineralization by electro-Fenton method (Kayan et al.,
2010).

These derivatives were further oxidized to quinoid structures (V–XI) and


formed carboxylic acids via ring opening. Finally, carboxylic acids were con-
verted to CO2 and H2O. In general, OH attacks C-H bonds nonselectively, but
is less reactive with aliphatic polyhalogenated compounds (Buxton, Greenstock,
Helman, & Ross, 1988, Liu et al., 2015). The relatively slow rate between OH
aliphatic polyhalogenated compounds can be attributed to the interference of
6 H. HE AND Z. ZHOU

halogen ions.
R ¡ H C OH ! R C H2 O (9)
R C O2 ! ROO ! products of degradation (10)
RH C OH ! RHðOHÞ (11)
C D C C OH !  C ¡ C ¡ OH (12)
ArH C OH ! ArHðOHÞ (13)

In the ferric system, Fe3C acts as a catalyst to decompose H2O2 to O2 and H2O,
and a steady-state concentration of Fe2C can be regenerated (Eqs. 14 and 15) (Sun
& Pignatello, 1993). Additionally, Fe2C and Fe3C in the system can act as coagu-
lants and lead to coagulation that further improves overall removal efficiency of
organic compounds.
Downloaded by [University of Florida] at 00:51 19 December 2017

Fe3 C C H2 O2 $ Fe ¡ O2 H2 C $ Fe2 C C HO2  (14)


3C  2C C
Fe C HO2 ! Fe C H C O2 (15)

Loss of iron is still an issue due to the precipitation of Fe (III) in electro-Fenton


oxidation. Fe3C generated from chain Fenton reactions may undergo hydration
(Eqs. 16–20), and the ferric oxyhydroxide species would precipitates out at high Fe
(II) concentrations in oxygen saturated solutions. Iron sludge will be eventually
produced in Fenton processes though the use of electricity can reduce its quantity
(Qiang, Chang, & Huang, 2003). Iron deposit (Fe(OH)3) could occur at pH over
2.5 and become especially important at pH 4 resulting in the dramatically
decreased current efficiency (Chou, Huang, Lee, Huang, & Huang, 1999, Duester-
berg, Mylon, & Waite, 2008; Kishimoto, Kitamura, Kato, & Otsu, 2013). Once the
iron deposit is formed, a pH value lower than 2 is necessary to re-dissolve it
completely due to its highly heterogeneous nature (Qiang et al., 2003). Ideally a
very acidic condition is favorable for Fenton methods, but such low pH is rarely
used in practices. Due to the neutral or alkaline pH in the water and wastewater
and highly concentrated iron species in use, the precipitation of ferric hydroxide
(Fe(OH)3) results in a major loss of iron.

Fe3 C C OH ¡ $ FeðOHÞ2 C (16)


2C ¡ C
FeðOHÞ C OH $ FeðOHÞ2 (17)
C ¡
FeðOHÞ2 C OH $ FeðOHÞ3 (18)
FeðOHÞ3 C FeðOHÞ3 $ precip (19)
FeðOHÞ3 C precip ! precip (20)

It should be noted that the hydroxyl radical theory has been controversial and
several studies in the literature suggested the generation of ferryl ion (FeO2C) with
the reactions between H2O2 and Fe2C (Eqs. 21 and 22), which act as reactive
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 7

species in the Fenton system (Bray & Gorin, 1932; Deguillaume, Leriche, &
Chaumerliac, 2005). New studies suggested that the Fenton mechanisms may
include hydroxyl radical theory or ferryl ion theory (Gozzo, 2001; Hug & Leupin,
2003). Operating parameters, such as pH, solvents, and the contact between a
metal ion and hydrogen peroxide, may influence the formation of oxidants and
also contribute to the formation of a high valence species of iron (FeO2C).

Fe2 C C H2 O2 ! FeO2 C C H2 O (21)


2C 2C
FeO C H2 O2 ! Fe C H2 O C O2 (22)

3. Kinetics modelling
Kinetic modeling is a crucial approach to describe and predict the performance of a
Downloaded by [University of Florida] at 00:51 19 December 2017

system. Many studies have been done on the kinetics of organic degradation in an
electro-Fenton system to elucidate the reaction mechanism and evaluate the pro-
cess performance. Various rate constants involved in Fenton reactions described
by pseudo-second-order kinetics are summarized in Table 2.
Reactions are associated with HC and OH¡, and thereby kinetic constants are
highly dependent on pH conditions. Generally rate constants increase with the pH
value due to the enhanced redox cycling of iron at higher pH (Skoumal et al.,
2009). Increased redox cycling of iron species can also explain the increase in the

Table 2. Model and rate constants for electro-Fenton reactions.


Rate constant (M¡1s¡1)

Reaction pH 2.5 pH 3.0 pH 4.0 pH 5.0 References

1 Fe2 C C H2 O2 ! Fe3 C 55 55 55 (Duesterberg and Waite


C OH C H2 O 2006, Duesterberg, Mylon
and Waite 2008)
63 1.2 £ 102 5.7 £ 102 Wang, Zheng, Zhang and
Wang (2016)
41.7 41.7 Duesterberg and Waite
(2006)
3C 2C ¡3 ¡3
2 Fe C H2 O2 ! Fe C 2.0 £ 10 2.0 £ 10 2.5 £ 10¡3 Duesterberg et al. (2008)
HO2  C H C
3 Fe3 C C O2  ¡ 6 HO2  1.24 £ 106 1.76 £ 106 7.70 £ 106 (Duesterberg and Waite
! Fe2 C C O2 C H C 7.82 £ 105 6.84 £ 106 3.1 £ 107 2006, Duesterberg et al.,
2008, Kwan and Voelker
2002)
4 Fe2 C C OH ! Fe3 C C 3.2 £ 108 3.2 £ 108 3.2 £ 108 (Duesterberg and Waite
OH ¡ 2006, Duesterberg et al.,
2008)
5 Fe2 C C O2  ¡ 6 HO2  1.24 £ 106 1.34 £ 106 2.40 £ 106 Duesterberg et al. (2008)
! Fe3 C C H2 O2
6 H2 O2 C OH ! O2  ¡ 6 3.3 £ 107 3.3 £ 107 3.3 £ 107 3.3 £ 107 (Buxton, Greenstock, Helman
HO2  C H2 O and Ross 1988, Kwan and
Voelker (2002)
2.7 £ 107 Buxton et al. (1988)
7 OH C O2  ¡ 6 HO2  ! 6 £ 109 Buxton et al. (1988)
H2 O C O2 7.11 £ 109 7.15 £ 109 7.51 £ 109 Duesterberg et al. (2008)
8 OH C OH ! H2 O2 5. £ 109 Buxton et al. (1988)
5.2 £ 109 5.2 £ 109 5.2 £ 109 Duesterberg et al. (2008)
8 H. HE AND Z. ZHOU

rate constant of the initiation reaction at higher pH (reaction 1) (Anotai, Lu, &
Chewpreecha, 2006). The majority of OH produced by reaction 1 is converted to
superoxide radicals (O2¡/HO2) via reaction 6. Superoxide radicals can be oxi-
dants for Fe2C and reductants for Fe3C. From Table 1, the reduction of Fe3C by
O2¡/HO2 (reaction 3), a propagating reaction, turns faster than the oxidation of
Fe2C by O2¡/HO2 (reaction 5) that acts as the terminating reaction with the
increase in pH, leading to a higher net iron concentration and an increased oxidiz-
ing capacity of the system (Duesterberg et al., 2008).
Reaction 4 and 6 prove the potential ability of Fenton’s reagents (Fe2C and H2O2)
to compete with target molecules for OH under certain conditions, and such ability
associates with the ratios of [Fe2C]0/[H2O2]0 (the initial concentration of Fe2C versus
that of H2O2) (Anotai et al., 2006). If the ratio is larger than one, implying that excess
Fe2C is supplied in the system, reaction 4 is predominant, while H2O2 becomes a
Downloaded by [University of Florida] at 00:51 19 December 2017

dominant OH scavenger at a ratio less than 1. HO2 generated from the H2O2-scav-
enging reaction (reaction 6) also attacks and oxidizes organic molecules, which pro-
vides an alternative reason for the promoted oxidation rates of contaminants with
the increase in H2O2 concentrations (Anotai et al., 2006).
In contrast to conventional Fenton systems, catalyst Fe2C can be regenerated
mainly via reaction 3 in an electro-Fenton system. The reported rate constants of
reaction 3 at all given pH conditions are way higher than rate constants of the initi-
ation reaction (reaction 1), implying that the regeneration of catalyst occurs so fast
that the Fenton chain reaction proceeds simultaneously if H2O2 is still available in
the system. Therefore, the concentration of OH can be considered independent
on initial Fe2C concentrations, and the degradation of organics by electro-Fenton
methods approaches zero-order with respect to [Fe2C]0 (Anotai et al., 2006; Zazo,
Casas, Mohedano, Gilarranz, & Rodriguez, 2005). However, it should be noted
that this condition is restricted to low concentrations of Fenton’s reagents and con-
taminants. During water and wastewater treatment, a majority of iron escapes
from the system via precipitation (as discussed in Section 2), and the iron hardly
exists in the redox cycle.

4. Impacting factors
4.1. Fe2C/Fe3C concentration
The availability of Fe2C is an essential prerequisite in an electro-Fenton process. In
some electro-Fenton systems, Fe2C is electro-regenerated from Fe3C. Therefore,
the concentration of Fe2C or Fe3C has a significant impact on the overall treatment
efficiency. Usually the efficiency and degradation rate of an electro-Fenton process
increase under a high concentration of iron species, as Fe2C promotes the genera-
tion of the hydroxyl radicals. Previous studies showed that Fe2C enhanced the oxi-
dizing power of H2O2 to destroy large molecules, such as dyestuffs, in real dyeing
wastewater (Wang, Chou, Chung, & Kuo, 2010), and COD removal efficiency
increased from 19.8% to 43.1% with the presence of a Fe2C concentration of
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 9

0.33 mM. Panizza and Cerisola found that the existence of Fe2C accelerated the
oxidation rate of alizarin red, and the final mineralization reached 93% with a Fe2C
concentration of 1.0 mM (Panizza & Cerisola, 2009). On the other hand, a higher
concentration of iron species enhances the ionic strength of the solution and there-
fore improves the current efficiency in an electro-Fenton system. A previous study
showed that the current efficiency was nearly 90% for a Fe3C concentration of
7000 mg/L, but only 39% for 1000 mg/L when pH was between 1.0 and 2.0. How-
ever, when pH increased from 2.0 to 2.5, a 75% drop in the current efficiency
occurred in the trial of 7000 mg/L ferric concentration while a 30% decrease was
observed in the 1000 mg/L trial (Chou et al., 1999). This observation revealed the
adverse effect of excess iron species. High concentrations of Fe2C or Fe3C lead to
the formation of iron precipitation at higher pH values (>2.5) and decrease the
current efficiency. An enormous increase of iron species also diminishes the con-
Downloaded by [University of Florida] at 00:51 19 December 2017

centration of OH (Eq. 5), increases the effluent electrical conductivity (Deng &
Englehardt, 2006), and produces iron salt that contributes to a problematic high
effluent concentration of suspended solids (Babuponnusami & Muthukumar,
2014). Therefore, the optimum concentration of ferrous or ferric ions should be
evaluated in laboratory-scale studies to maximize electro-Fenton efficiency before
large-scale applications.

4.2. H2O2 concentration


The concentration of H2O2 plays a significant role in the efficiency of an electro-Fen-
ton reaction, as it directly influences the theoretic peak amount of produced OH
(Deng & Englehardt, 2006). An increase in the H2O2 dosage usually improves the
overall efficiency (Lin, Lin, & Leu, 1999; Moon, Ezuka, Maruyama, Osakada, &
Yamamoto, 1993; Ratanatamskul, Masomboon, & Lu, 2011), which is due to the
increase of OH (Eq. 1). Many studies have demonstrated this correlation. For exam-
ple, increase of H2O2 dosage from 0 mg/L to 1600 mg/L significantly increased both
overall COD removal efficiency and oxidation efficiency (Gogate & Pandit, 2004),
which was consistent with the results of another study of phenol oxidation (Gumus
& Akbal, 2016). However, large quantities of H2O2 reduce removal efficiency by
scavenging generated OH (Eqs. 2 and 23) or recombining OH (Eq. 24).

HO2  C OH ! H2 O C O2 (23)


2OH ! H2 O2 (24)

Additionally, unused portion of H2O2 with reductive ability can consume chem-
ical oxidant during COD analysis thus leads to overestimation of COD values, and
the extent of error is proportional to H2O2 concentration (Kang, Cho, & Hwang,
1999). Therefore, the excess portion of H2O2 is likely to cause reduction of organ-
ics, thus Lin and Lo reported that 1 mg/L H2O2 could contribute 0.27 mg/L COD
in the treatment of desizing water (Lin & Lo, 1997). Lee et al. reported that the
10 H. HE AND Z. ZHOU

average overestimation of COD in livestock wastewater was 0.52 mg COD/mg


H2O2 in the COD range of 0 to 400 mg/L (Lee, Lee, Kim, Sohn, & Lee, 2011). Cor-
rection of H2O2 interference has been analyzed in previous studies (Kang et al.,
1999; Lee et al., 2011).
Besides, the excess amount of H2O2 is toxic to many organisms and dramatically
decreases overall efficiency in those cases where the Fenton technology functions as
a pretreatment to biological processes (Ito et al., 1998). A suitable H2O2 concentra-
tion needs to be determined to maximize the efficiency in an electro-Fenton system
while minimizing concomitant drawbacks.

4.3. Initial concentration of pollutants


The initial concentration of pollutants affects the overall efficiency of electro-Fenton
processes. Dosages of Fenton’s reagents need to be determined based on the initial
Downloaded by [University of Florida] at 00:51 19 December 2017

concentration of pollutants for treatment. Usually, low initial concentrations are pre-
ferred (Benitez, Acero, Real, Rubio, & Leal, 2001; Hou et al., 2016; Kwon, Lee, Kang,
& Yoon, 1999; Rosales, Pazos, Longo, & Sanroman, 2009) and dilution is an essential
step to ensure a good removal efficiency in electro-Fenton systems (Gogate & Pandit,
2004). A previous study indicated that that higher initial concentrations of dyes con-
tributed to the production of reactive intermediates, which reduced the amount of
OH to attack target pollutant and therefore decreased the overall efficiency (Rosales
et al., 2009). However, in large-scale industrial wastewater treatment, dilution creates
a large quantity of wastewater and substantially increases treatment costs. Pollutants
need to be properly diluted to properly balance the benefits and costs for cost-effi-
cient removal of organic pollutants in electro-Fenton systems.
In spite of possible low removal efficiency under high concentrations, the amount
of removed mass of pollutants may still be high under higher initial concentrations
of pollutants (Jasmann, Borch, Sale, & Blotevogel, 2016). As previously reported, the
amount of removed COD increased from 613 mg/L to 1124 mg/L when initial COD
increased from 1000 mg/L to 3000 mg/L, although COD removal efficiency
decreased from 61.3% to 37.5% (Zhang, Choi, & Huang, 2005). However, initial con-
centrations may not always have significant impacts. In another study, the initial
amount of the herbicides did not affect the removal of atrazine and metolachlor
(Pratap & Lemley, 1998). A possible reason is that the oxidation performance was
limited by the generation rate of OH and its side reactions with other radical scav-
engers. The rate constants of reactions between herbicides and OH were near the
diffusion-controlled limit (1010 M¡1s¡1) (Pratap & Lemley, 1998), and the initial
availability of herbicides did not have a great impact on the oxidation rates.

4.4. pH
The value of pH is one of the main factors that affect the performance in an elec-
tro-Fenton system. Acidic medium is preferred as basic pH adversely affects the
overall treatment efficiency. The optimum pH value was reported to be around
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 11

three in several studies regardless of the types of target contaminants (Hammami,


Oturan, Bellakhal, Dachraoui, & Oturan, 2007; Ma et al., 2016; Nidheesh,
Gandhimathi, Velmathi, & Sanjini, 2014). In another study, a pH of two was
reported as the optimal value (Ting, Lu, & Huang, 2009). In addition, some studies
indicated that H2O2 generation was maximized at neutral pH (Garcia et al., 2013;
Liu et al., 2015). The inconsistency among these studies could be attributed to the
variability of target pollutants and experimental conditions. Some previous studies
showed that the Fenton process was not robust in degrading contaminants at high
pH values due to the oxidation, hydrolysis and precipitation of iron ions (Brillas &
Casado, 2002; De Laat, Truong Le, & Legube, 2004; Ratanatamskul et al., 2011), in
which iron species possibly exist in the form of Fe3C at nanomolar levels at a pH of
4.4 (Stumm & Morgan, 1985). Also, H2O2 does not generate OH under basic con-
ditions (Rivas, Beltran, Gimeno, & Carvalho, 2003), which could be caused by
Downloaded by [University of Florida] at 00:51 19 December 2017

decomposition of H2O2 by OH¡ (Eq. 25) (Liu et al., 2015). At acidic solutions,
more protons are available in the solution that can promote the conversion from
dissolved oxygen to H2O2 (Eq. 6). Sedlak and Andren explained that OH produc-
tion was high in the pH range of 2–4 due to reactions involving the organometallic
complex where H2O2 is regenerated (Sedlak & Andren, 1991).

H2 O2 C 2OH ¡ ! O2 C 2H2 O C 2e ¡ (25)

However, very low pH values also adversely affect system performance for the
following reasons. First, at low pH conditions, hydrogen evolution is also pro-
moted, which reduces the number of available protons to generate H2O2 (Eqs. 6
and 26) and promote the decomposition of H2O2 (Eq. 27) (Wang et al., 2010;
Wang, Hu, Chou, & Kuo, 2008). Second, iron complex species (Fe(H2O)6)2C) exist
in the low pH environment, which reacts slowly with H2O2 and reduces the gener-
ation rate of OH (Kavitha & Palanivelu, 2005). The concentration of Fe2C contrib-
utes to the formation of Fe(H2O)62C as well. Finally, in the presence of high
concentrated protons, H2O2 trends to form stable oxonium ion (H3O2C), which
reduces the reactivity of H2O2 with Fe2C (Kavitha & Palanivelu, 2005; Kwon et al.,
1999).

2H C C 2e ¡ ! H2 (26)
C ¡
H2 O2 C 2H C 2e ! 2H2 O (27)

On the other hand, due to the formation of Fe(OH)n-type structures at higher


pH conditions (n D 2 or 3), an increase in pH may lead to electrocoagulation that
helps get rid of pollutants via electrostatic attraction and complexation (Mollah,
Schennach, Parga, & Cocke, 2001). Besides, an increase in pH helps decrease
the oxidation potential of OH and therefore reduces the threshold of OH genera-
tion (Kavitha & Palanivelu, 2005). The oxidation potential for the redox couple
OHaq/H2O was reported to be reduced from 2.59 V at pH 0 to 1.64 V at pH 14.0
12 H. HE AND Z. ZHOU

(Bossmann et al., 1998). As the overall treatment efficiency of electro-Fenton is


diminished at both high and low pH values, pH should be optimized for improved
performance of electro-Fenton systems.
In recent years, it is remarkable that in situ generation of Fenton’s reagents (Gao
et al., 2015; Liu et al., 2015; Vermilyea & Voelker, 2009; Wang et al., 2013; Wang,
Liu, Yuan, & Li, 2014) and biological production of electrons (Feng, Li, Mai, & Li,
2010) have allowed Fenton being operated at a neutral pH, which solved the issue
of feeding acids regularly to keep iron species dissolved.
It should be noted that electro-Fenton efficiency can be affected by buffer solu-
tions that are used to maintain stable pH in experiments. In a treatment study of
p-hydroxy phenylacetic with electro-Fenton, acetic acid buffer led to the optimal
oxidation performance while phosphoric or sulfuric acid buffer reduced removal
efficiency (Benitez et al., 2001), which is probably due to the formation of iron
Downloaded by [University of Florida] at 00:51 19 December 2017

complexes with the buffer that decreased the reactivity of Fe2C (Benitez et al.,
2001; Pignatello, 1992). Reaction buffers increase additional treatment costs, and
therefore the selection and utilization of buffers need to be balanced with treatment
efficiency (Babuponnusami & Muthukumar, 2014).

4.5. Electrolyte
An electrolyte is necessary as it improves the conductivity of solutions for efficient
electro-Fenton reactions. Different types of electrolytes have been used in electro-
Fenton systems. Sodium sulfate (Na2SO4) is the most commonly used electrolyte
in electro-Fenton processes for its high ionic strength and low interference in aque-
ous solutions. Other electrolytes such as KCl, NaCl, NaNO3, and NaClO4, have
been adopted as well. In the study by Ghoneim et al., the efficiencies of KCl, NaCl,
and Na2SO4 at the same concentration (0.05 M) was compared and SO42¡ gave
the maximum decoloration rate of sunset yellow (Ghoneim, El-Desoky, & Zidan,
2011). It’s also reported that the removal rate of Orange II followed the order of
ClO42¡ > Cl¡ > SO42¡ (Daneshvar, Aber, Vatanpour, & Rasoulifard, 2008). H2O2
accumulation was faster in the presence of NO3¡ than that of SO42¡ and minerali-
zation efficiency was better for ClO42¡ media than SO42¡ media. The difference
can be attributed to the formation of chloro and sulfato-complexes that decreased
the concentration of free Fe2C while no complexes were noted in the presence of
ClO42¡ or NO3¡ (De Laat et al., 2004; Qiang, Chang, & Huang, 2002).
When chloro-salts are used as the electrolyte, it is likely that the presence of
¡
Cl might promote the degradation efficiency and shorten the reaction time by
producing chlorine (Cl2(aq)) or hypochlorous acid (HClO), which are also strong
oxidizing agents in aqueous solutions (Eqs. 28 and 29) (Babu, Venkatesan,
Kanimozhi, & Basha, 2009; Shan et al., 2016). In the study of acid orange 7 removal
(Fernandes, Morao, Magrinho, Lopes, & Goncalves, 2004), the electrolyte KCl
resulted in a better removal performance of Acid Orange 7 on BDD anodes than
NaSO4 due to extra oxidizing agents (hypochlorite) electrogenerated from Cl¡. In
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 13

addition, it has been discovered that Cl¡ can strengthen H2O2 accumulation (Da
Pozzo, Petrucci, & Merli, 2008). Since the SCE reference electrode was filled with
KCl solution, the diffusion of Cl¡ from the SCE to catholyte could promote
cathodic H2O2 generation (Qiang et al., 2002).

2Cl ¡ ! 2e ¡ C Cl2 (28)


C ¡
Cl2 C H2 O ! H C Cl C HClO (29)

Even though the electro-generated chlorine species, such as HClO, ClO¡, Cl,
and Cl2, can participate in the oxidation process and enhance the overall electro-
Fenton efficiency, it should be noted that such a mixture is also known to form
many organochlorinated species as intermediates. The formation of unwanted
compounds are mainly due to the presence of OCl¡ and can only be avoided if
Downloaded by [University of Florida] at 00:51 19 December 2017

organics are completely mineralized (Fernandes et al., 2004). In the study of the
electrochemical oxidation of phenol in the presence of NaCl (Comninellis &
Nerini, 1995), nonvolatile organochlorinated compounds (chlorinated aliphatic
acids) were formed in the electrolyte and finally oxidized to volatile chlorinated
compounds (CHCl3). Some chlorinated organics, such as 4-chlorophenol, 2,4-
dichlorophenol, and 2,4,6-trichlorophenol, can be even more hazardous than raw
phenol, requiring additional treatments and significant cost (Ca~ nizares,
Garcıa-Gomez, Saez, & Rodrigo, 2004).
Cl¡ has been well demonstrated as a scavenger of OH (Eqs. 30 and 31) (Liao,
Kang, & Wu, 2001) and OH yield decreases with the increase in NaCl concentra-
tion (Liao et al., 2001; Shan et al., 2016). Chloride anions can be oxidized to form
chloride atoms (Cl), which forms dichloride anion radicals (Cl2¡) subsequently.
OH is more reactive than Cl and Cl2¡ (De Laat et al., 2004; Shan et al., 2016),
and therefore the steady-state concentration of OH is much lower than that of Cl
and Cl2¡. In another case, the oxidation of tetramethylbenzidine increased by
50% at 0.1 M NaCl (Shan et al., 2016). This finding supports the theory that Cl¡
could enhance the removal efficiency, yet in these systems, chloride radicals instead
of OH may be the dominant oxidizing agents for contaminant removal.

OH C Cl ¡ !  HOCl ¡ (30)


 ¡ C 
HOCl C H ! Cl C H2 O (31)

The concentrations of electrolytes also play an important role in electro-Fenton


efficiency. Low concentrations of electrolytes might not provide enough conductiv-
ity while high concentrations of electrolytes could cause electrode corrosion and
shorten the reactivity of Fenton species (Ruiz et al., 2013). Researchers reported
that a high current density was observed in solutions with a high Na2SO4 concen-
tration, which was corresponding to the increased H2O2 production rate, enhanced
efficiency, and reduced energy consumption (Choi, Lee, Shin, & Yang, 2010; Zhou,
Yu, Lei, & Barton, 2007). The methyl red removal efficiency increased from 56.8%
14 H. HE AND Z. ZHOU

at 0.05 M Na2SO4 to near 80% at 0.1 M Na2SO4, which was observed as the optimal
concentration. Further increase of the Na2SO4 concentration to 0.2 M significantly
dropped removal efficiency, which may be due to the consumption of OH by
SO42¡ (Eqs. 32 and 33).

¡
OH C SO4 2 ! OH ¡ C SO4  ¡ (32)
¡
SO4  ¡ C e ¡ ! SO4 2 (33)

However, the correlation between electrolytes and electro-Fenton efficiency was


not always obvious. In one study, an increase in the Na2SO4 concentration from
0.05 M to 0.1 M did not affect the generation of H2O2 in reactions (Daneshvar
et al., 2008). In another study, formaldehyde removal was not influenced by
Downloaded by [University of Florida] at 00:51 19 December 2017

Na2SO4 concentration in the range of 0–0.5 M (Do & Chen, 1994). Other chemi-
cals in wastewater may could be utilized as supporting electrolytes for on-site
H2O2 production (Qiang et al., 2002).

4.6. Oxygen sparging rate


Oxygen sparging rate is a major factor that limits electro-Fenton efficiency. A
greater oxygen sparging rate can raise the dissolved oxygen (DO) level and mass
transfer rate in the system, and therefore promote H2O2 production that the
removal efficiency depends on. Wang et al. reported that the increase in oxygen
sparging rate from 50 cm3/min to 250 cm3/min led to around 30% increase in the
COD removal percentage under a constant current density of 3.2 mA/cm2 (Wang
et al., 2010). It should be noted that the amount of generated H2O2 does not

increase linearly with oxygen flow rates (Ozcan, Şahin, Savaş Koparal, & Oturan,

2008). After DO is saturated (8.26 mg/L at 25 C (Brezonik & Arnold, 2011)), fur-
ther increase of oxygen sparging rate does not increase DO concentration in the
system and H2O2 generation trends to be stable. In fact, further benefits from
excess increase of the oxygen sparging rate is limited. According to study by Wang
et al., color removal efficiency of the real dyeing wastewater increased from 55.3%
to 70.0% when the oxygen sparging rate increased from 100 cm3/min to 300 cm3/
min, but remained stable at 70.6% under oxygen sparging rate of 400 cm3/min
(Wang et al., 2008).

4.7. Feeding mode


The feeding mode of Fenton’s reagents has a strong correlation with organic
removal efficiency and operational costs of electro-Fenton processes. The addition
of H2O2 before or during the reaction influences H2O2/Fe2C, H2O2/COD, and sub-
sequent COD removal efficiency (Primo, Rivero, & Ortiz, 2008). Continuous or
stepwise addition help promote the oxidizing capacity and removal efficiency
(Duesterberg & Waite, 2006; Primo et al., 2008; Zhang et al., 2005; Zhang, Fei,
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 15

Zhang, & Tang, 2007). Adding H2O2 multiples time with small quantities could
leave much of OH available to attack and direct oxidize the target substrate
(Duesterberg & Waite, 2006). One-step addition can result in self-decomposition
of H2O2 (Eq. 34) due to highly localized concentrations at the feeding point and
consumption of OH by H2O2 (Eq. 2) (Deng & Englehardt, 2006).

2H2 O2 ! 2H2 O C O2 (34)

Factors such as feeding dosage, time interval, and flow rate over which H2O2 is
injected can be determined based on an analysis of the kinetic model to reach the
maximum efficiency (Duesterberg & Waite, 2006). Primo et al. investigated the
effects of H2O2 feeding mode, and found that COD removal efficiency increased
from 78% at one H2O2 dose to 86% at four H2O2 doses ([H2O2] D 10,000 mg/L)
Downloaded by [University of Florida] at 00:51 19 December 2017

(Primo et al., 2008). This result is consistent with the result in another study, in
which COD removal efficiency increased with the number of H2O2 treatments and
reached the peak at a continuous feeding mode (Zhang, Choi, & Huang, 2006).
Nevertheless, according to Wang et al., two or three-step addition of H2O2 did not
significantly change the COD removal rate (56% for two-step addition and 58%
for three-step addition), and therefore the twice dosing schedule was preferred to
achieve removal and cost efficiency (Wang, Chen, Gu, & Wang, 2009).

4.8. Current density


The performance of electro-Fenton systems is considerably affected by applied cur-
rent density, which is the driving force of electron transfer and therefore corre-
sponds to the rate of H2O2 generation. High current density enlarges the amount
of OH produced in the solution and physisorbed on electrodes. Also, a higher cur-
rent density leads to a faster regeneration of Fe2C (Eq. 35) and increases the effi-
ciency of Fenton chain reactions (Zhang et al., 2007). As was reported by Zhang
et al., COD removal efficiency improved nearly 40% when current raised from
100 mA to 250 mA (Zhang et al., 2006). Similarly, an increase of 20.9% in TOC
mineralization efficiency was observed when the current density increased from
10 mA/cm2 to 24 mA/cm2 (Hou et al., 2016). The enhanced degradation rate at
higher current density results in a reduced reaction time, but it should be noted
that the time-saving benefit becomes less significant if the applied current density
is too high (Choi et al., 2010; Skoumal et al., 2009).

Fe3 C C e ¡ ! Fe2 C (35)

Current efficiency may decline under high applied current density (Brillas et al.,
2004; Hou et al., 2016; Zhang et al., 2006) since higher current density also acceler-
ates the occurrence of side reactions, such as anodic oxygen discharge (Eq. 36),
cathodic hydrogen evolution (Eq. 26), and parasite reactions of OH (Eq. 37). In a
16 H. HE AND Z. ZHOU

study on leachate treatment, COD removal efficiency was 89.2% at 250 mA, but
dropped to 79.3% at 300 mA (Zhang et al., 2006). In the mineralization of desme-
tryne, TOC removal efficiency improved 42% when applied current increased
from 100 mA to 300 mA, but only improved 10% after current increased to
450 mA (Borras et al., 2011). Meanwhile, the mineralization current efficiency
dropped from 13.7% at 100 mA to 7.1% at 450 mA (Borras et al., 2011). According
to Shin et al., the kinetic rate of electro-Fenton reached the peak when the current
density was 28.5 mA/cm2 (Shin et al., 2017). No further improvement was
observed with the increase in the current density due to possible reactions between
organic substrates and OH that reduced the oxidizing agent amount.

2H2 O ! 4H C C O2 C 4e ¡ (36)
2OH ! O2 C 2H C C 2e ¡ (37)
Downloaded by [University of Florida] at 00:51 19 December 2017

Additionally, the current control regime where the degradation increases line-
arly with applied current becomes shorter with higher current density, which
means the current control regime ends at a higher residual COD level with the
higher current density (Choi et al., 2010). Thus, current efficiency can be higher at
a lower current density because the current control region is longer where most of
the applied current is used to degrade target contaminants. When a system exceeds
the current control regime, the promotion brought by exceeding higher current
density is slight due to the limit of mass transport (Choi et al., 2010). In a previous
study, degradation rate was enhanced within 4 hr when current density increased
from 6.6 mA/cm2 to 33.3 mA/cm2, while a similar rate was obtained at a longer
time due to the generation of by-products, but when the current density further
increased from 33.3 mA/cm2 to 100 mA/cm2, degradation rate did not rise accord-
ingly because of the kinetic limitation of the decontamination process by the mass
transfer of organics towards electrodes and their diffusion in the solution (Skoumal
et al., 2009).
Higher current density also requires more energy consumption. Therefore, an
increase in current density brings the benefits of enhanced degradation rates and
less removal time and the drawbacks of increased energy consumption and
decreased current efficiency. It is necessary to adjust the applied current density
for a balance between the desired efficiency and energy costs.

4.9. Operating temperature


The temperature of a reaction system is a crucial parameter as it is significantly
correlated with reaction rates. An increase in the operating temperature likely
accelerates the reaction rate between H2O2 and Fe2C, thus boosts the OH genera-
tion and contaminants degradation. A temperature lower than 18.3 C can slow
down initial kinetics and negatively affect the reaction rate and removal efficiency
(Deng & Englehardt, 2006).
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 17

The different effects of temperature on electro-Fenton efficiencies may be


explained by decreased oxygen solubility in water (Ozcan € et al., 2008) and high
decomposition rate of H2O2 into inactive species under high temperatures (Eq. 34)

(Ozcan et al., 2008). In fact, a 10 C rise in temperature would accelerate the H2O2
decomposition by 2.3 times (Jones, Braithwaite, & Clark, 1999) and temperature
over 50 C contributes to the destabilization of flocs and significantly decrease
COD removal efficiency (Umar, Aziz, & Yusoff, 2010). Wang et al. observed that
the COD removal dropped from 75.2% to 68.1% as the temperature rose from
20 C to 40 C (Wang et al., 2010).
Temperature also affects the current efficiency in electro-Fenton processes
(Choi et al., 2010). In a study where three temperature values (5 C, 25 C, and
55 C) were used (Choi et al., 2010), the current control regime lasted longer under
higher temperatures, meaning a better current efficiency under higher tempera-
Downloaded by [University of Florida] at 00:51 19 December 2017

tures. When a system exceeded the current control regime, the authors observed
that even though the current efficiency of a higher temperature decreased more
sharply, it remained a higher level than that of lower temperatures at the same
COD concentration.
It should be noted that the influence of temperature is less significant than other
factors (Umar et al., 2010; Zhang et al., 2005) and ambient temperature has been
frequently used in many electro-Fenton systems with good efficiencies. If a temper-
ature over 40 C is likely to occur during exothermic reactions, cooling is recom-
mended to slow down the accelerated decomposition of H2O2 (Gogate & Pandit,
2004).

4.10. Electrode materials


A fundamental step in an electro-Fenton process is to select electrode materials.
Proper anode materials can prevent potential deterioration of electrodes, and high
oxygen overvoltage anode can efficiently produce hydroxyl radicals (Eq. 8) to
enhance treatment efficiency (Zhang et al., 2007).
Platinum (Pt) is potent for acting as an anode in electrochemical processes due
to its excellent conductivity and high stability (Nidheesh & Gandhimathi, 2012).
Previously used platinum anodes include Pt flake (Liu et al., 2007), Pt wire
(Wang et al., 2010), Pt sheet (Brillas et al., 2007; Daneshvar et al., 2008; Ham-
mami et al., 2007; Sires et al., 2007), Pt gauze (Ghoneim et al., 2011; Ozcan,
Sahin, Koparal, & Oturan, 2008), Pt grid (Oturan, Panizza, & Oturan, 2009;
Pimentel, Oturan, Dezotti, & Oturan, 2008), Pt mesh (Balci, Oturan, Oturan, &
Sires, 2009; Celebi, Oturan, Zazou, Hamdani, & Oturan, 2015), and Pt plate (Ede-
lahi et al., 2003). However, the cost-effectiveness of Pt is not good because of its
high price. To address this problem, platinized anodes, which is a technique of
coating or plating certain quantity of platinum on suitable metals, has been uti-
lized in many electro-Fenton processes and could achieve comparable results as
platinum anodes. For example, platinized titanium (Ti/Pt) (Brillas & Casado,
18 H. HE AND Z. ZHOU

2002; Casado, Fornaguera, & Galan, 2005) and Pt-coated steel electrodes (Choi
et al., 2010) have been used in early studies. Other noble metals, such as gold
(Au) and silver (Ag), can also function as anode electrodes, yet their high costs
make them unsuitable for practical applications.
Boron-doped diamond (BDD) is a suitable electrode material for electrochemi-
cal applications. BDD present a higher sensitivity and lower detection limit due to
its wide potential window and low background currents in aqueous solutions.
Over the years, BDD has been utilized as an anode (Brillas et al., 2007; Celebi et al.,
2015; Wang et al., 2013) or both anode and cathode (Garcia et al., 2013). Com-
pared with Pt electrodes, BDD helps produce a much greater quantity of OH and
can completely remove unsaturated and aromatic compounds (Pliego et al., 2015).
Previous studies showed that BDD is excellent in oxidizing phenol (Lee et al.,
2017), carboxylic acids (Brillas, Garcia-Segura, Skoumal, & Arias, 2010; Lin et al.,
Downloaded by [University of Florida] at 00:51 19 December 2017

2013), and dyes (Panizza & Cerisola, 2008). The performance of BDD electrode in
the degradation of herbicide 2,4-DP was compared with Pt electrode by monitor-
ing the degradation efficiency of 2-(2,4-dichlorophenoxy)-propionic acid, and the
results showed that BDD electrode yielded better efficiency due to its higher elec-
trochemical activity and more production of oxidizing agents (OH) on its surface
(Brillas et al., 2007). This conclusion agrees with other studies (Choi et al., 2010,
Brillas, Boye, Sires, Garrido, Rodrıguez, Arias, Cabot and Comninellis 2004). Other
types of anodes used in electro-Fenton processes include DSA (titanium rod coated
with RuO2/IrO2) (Brillas & Casado, 2002; Chou et al., 1999; Huang, Huang, Chang,
& Chen, 2008), carbon felts (Feng et al., 2010; Wang et al., 2014), graphite (de
Dios, Rosales, Fernandez-Fernandez, Pazos, & Sanroman, 2015; Nidheesh et al.,
2014) and carbon nanotube (CNT) (Gao et al., 2015; Wang et al., 2014).
Proper cathode materials can directly improve electro-Fenton efficiency by pro-
moting the productivity of H2O2. The dissolved oxygen either transforms into
H2O via a 4-electron oxygen reduction reaction (ORR) (Eq. 38) or H2O2 by a 2-
electron ORR (Eq. 6), and the ORR activities are dependent on cathode materials.
Suitable cathode nature can avoid 4-electron ORR to maximize the H2O2 genera-
tion and electro-Fenton efficiency. Pore numbers and structure on the surface of
cathode materials affect mass transfer rate and electrochemical reaction areas for
the H2O2 production (Wang et al., 2014).

O2 C 4e ¡ C 4H C ! 2H2 O (38)

Carbon-based materials are commonly used as cathodes for electro-Fenton


systems for their appropriate electrochemical properties towards DO reduc-
tion, low catalytic activity for H2O2 decomposition, and high overpotential for
hydrogen evolution (Daneshvar et al., 2008). Previously used carbon-based
cathode materials for H2O2 generation include carbon felts (Balci et al., 2009;
Brillas et al., 2007; Celebi et al., 2015; Edelahi et al., 2003; Hammami et al.,
2007; Oturan et al., 2009; Ozcan et al., 2008; Pimentel, Oturan, Dezotti and
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 19


Oturan 2008; Wang et al., 2014), carbon sponge (Ozcan et al., 2008), reticu-
lated vitreous carbon (Ghoneim et al., 2011), CNT (Gao et al., 2015), activated
carbon fiber (Wang et al., 2010), graphite rod (Liu et al., 2007, Rosales et al.,
2009), graphite felt (Daneshvar et al., 2008), graphite paper (de Dios et al.,
2015), and graphite plates (Nidheesh et al., 2014). Modifications of cathode
materials have also been applied to produce more active sites at cathode sur-
face to enhance the H2O2 generation. A graphite-polytetrafluoroethylene
(PTFE) cathode has been implemented to an electro-Fenton system, and its
high H2O2 generation rate and current efficiency have been demonstrated
(Zhou et al., 2007). In the study of Feng et al., a composition of CNT and
FeOOH was used as the cathode material, which can release Fe2C to catalyze
oxygen reduction (Eq. 1) and therefore increase electro-Fenton efficiency
(Feng et al., 2010). Other modified carbon cathodes that have been used previ-
Downloaded by [University of Florida] at 00:51 19 December 2017

ously include hierarchically porous carbon (Liu et al., 2015; Rivas, Beltran,
Frades, & Buxeda, 2001), porous CNT sponge (Wang et al., 2014), and carbon
aerogel (Wang et al., 2013).

4.11. Distance between electrodes


Distance between electrodes affects the overall efficiency of electro-Fenton pro-
cesses as it governs the system potential and energy consumption. A reduction in
the interelectrode distance leads to a reduced resistance between electrodes and a
higher current flow. The traveling time of participating ions in electro-Fenton reac-
tions increases with the increase in the distance of electrodes, and results in a
higher duration electrolysis and lower removal efficiency (Verma, Khandegar, &
Saroha, 2013). In electro-Fenton processes, Fe2C can be electro-regenerated from
Fe3C on the cathode (Eq. 35), which is controlled by either electron transfer
between Fe3C and cathode or mass transfer of Fe3C across cathode-solution inter-
face (Qiang et al., 2003). Longer distance between electrodes contributes to the lim-
iting mass transfer of Fe3C to cathode surface (Zhang et al., 2006), and therefore
inhibits the catalyst regeneration and electro-Fenton efficiency. On the other hand,
the increase in the distance of electrodes gives a significant raise to energy expenses
(Atmaca, 2009). A previous study of leachate treatment by electro-Fenton showed
that COD removal efficiency dropped from 80.8% to 71.8% when the interelec-
trode distance increased from 2.1 cm to 2.8 cm (Zhang et al., 2006). However, the
COD removal efficiency for an interelectrode distance of 0.7 cm was even lower
than that for 1.3 cm (73.6% and 80.4%, respectively). This is because that the elec-
tro-regenerated Fe2C could be easily oxidized to Fe3C at the anode (Eq. 39) when
electrodes were placed too close. The optimal range of the interelectrode distance
was determined to be 1.3–2.1 cm.

Fe2 C ! Fe3 C C e ¡ (39)


20 H. HE AND Z. ZHOU

5. Comparison of performance, cost and energy efficiency


Existing literature showed a high performance of electro-Fenton process compared
with anodic oxidation, and conventional Fenton processes on the treatment and
mineralization of organic pollutants. Electro-Fenton process, as an indirect electro-
chemical approach, is generally considered more productive than direct anodic
oxidation with fewer problems of electrode fouling or corrosion (Martinez-Huitle
& Ferro, 2006). OH destroys organics more rapidly, thus making electro-Fenton
processes more efficient than anodic oxidation. In an early comparison of anodic
oxidation and electro-Fenton for the degradation of a 1000 mg/L aniline solution
(Brillas & Casado, 2002), electro-Fenton showed significantly higher oxidation
capacity with shorter electrolysis time and better energy efficiency than anodic oxi-
dation (Table 3).
Downloaded by [University of Florida] at 00:51 19 December 2017

Electro-Fenton processes are also more efficient than conventional Fenton


methods mainly due to its significant advantages of the electrolytic generation/
regeneration of Fenton’s reagents that allows for a high oxidation efficiency.
According to previous studies, electro-Fenton process competes the anodic oxida-
tion or conventional Fenton with respect to the degradation of dye (Bouzayani,
Meijide, Pazos, Elaoud, & Sanroman, 2017; Rosales et al., 2009), herbicides (Brillas
et al., 2004, 2007), drug ketoprofen (Feng, Oturan, van Hullebusch, Esposito, &
Oturan, 2014), PFOA (Liu et al., 2015), phenol (Gumus & Akbal, 2016), and dye-
reactive black B (Huang et al., 2008).
Energy consumption is the major concern for electrochemical wastewater treat-
ment. When additional assistance of electricity is used to enhance the treatment
efficiency, energy input and cost must be taken into consideration (Wang et al.,
2016). The calculations of input electric energy are shown in Eq. (40), where W is
the energy used per unit volume; Q is the specific electrical charge passed per unit
volume (kAh/m3), and V is the volume of treated water or wastewater (Wang
et al., 2016). Usually a higher energy efficiency occurs at the early stage of an elec-
trochemical treatment. With the proceeding of the treatment, the occurrence of
competing reactions and consumption of active sites on the electrode surface result
in more energy investment and cost.

WDQ  V (40)

Table 3. A comparison between anodic oxidation and electro-Fenton for the degradation of a
1000 mg/L aniline solution with and a carbon-PTFE O2-fed cathode (Brillas & Casado, 2002).
Solution Applied Electrolysis TOC Energy cost
Process Anode Volume (L) current (A) time (h) removal (%) (kWh/m3)

Anodic oxidation Ti/Pt 30 20 6 18 502


Electro-Fenton Ti/Pt 30 5 4 43 11
Electro-Fenton Ti/Pt 30 10 3 44 29
Electro-Fenton Ti/Pt 30 20 2 61 45
Electro-Fenton DSA 30 20 2 63 40
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 21

Among electrochemical treatments, electro-Fenton has been reported to con-


sume less energy per mole contaminants degraded, which can be attributed to a
lower required potential (¡0.4 V vs. SHE) for OH production. In the study of
PFOA degradation (Niu, Lin, Xu, Wu, & Li, 2012; Zhuo, Deng, Yang, Huang, &
Yu, 2011), the current efficiency for electro-Fenton process was one order of mag-
nitude higher than that reported for other electrochemical oxidation methods. The
energy consumption for electro-Fenton process was 0.42 kWh/(g PFOA), com-
pared to 1.13–12.5 kWh/(g PFOA) for photocatalysis and photolysis (Li et al.,
2012; Qu, Zhang, Li, Chen, & Zhou, 2010; Wang et al., 2010). These results sug-
gested that electro-Fenton process was an energy effective method to degrade
PFOA. Flow-through electro-Fenton shows an even better efficiency and efficacy
than conventional electro-Fenton. In addition, the use of flow-through configura-
tion (Gao et al., 2015), continuous aeration (Mousset, Wang, & Lefebvre, 2016),
Downloaded by [University of Florida] at 00:51 19 December 2017

and modified electrodes (Liu et al., 2017; Tian, Olajuyin, Mu, Yang, & Xing, 2016)
can further reduce the energy consumption of electro-Fenton processes.

6. Challenges and future perspective


Although Electro-Fenton process is a promising technique for water and wastewa-
ter treatment, there are still challenges. The low productivity of H2O2 is one chal-
lenge for electro-Fenton processes due to the low solubility of O2 at ambient
conditions. Electrode surface pretreatment technologies and development in cath-
ode materials can improve OH production and overall performance. Another
issue with electro-Fenton is pH control that requires large amount of chemicals.
As the efficiency of electro-Fenton method depends on pH, it is necessary to firstly
reduce the pH and subsequently neutralize the solution to eliminate the potential
pollution of low pH (Hammami et al., 2007). Iron sludge (Fe(OH)3) production is
still a problem in electro-Fenton process due to the low solubility of ferric iron.
Even though sludge can be reduced to Fe2C electrochemically, such a process
requires treatment under a pH of 2 (Chou et al., 1999), which is often impractical
in treatment systems and also raises the issue of pH neutralization.
Other electro-Fenton-based processes, such as photo-electro-Fenton (Casado
et al., 2005; Ding, Ai, & Zhang, 2012; El-Ghenymy et al., 2013; Ruiz, Arias, Brillas,
Hernandez-Ramırez, & Peralta-Hernandez, 2011; Ting, Lu, & Huang, 2008; Ver-
milyea & Voelker, 2009) and sonoelectro-Fenton are being developed to address
the issues of pH control and H2O2 production. The following trend was reported
in the case of degradation efficiency: photo-electro-Fenton > sono-electro-Fenton
> electro-Fenton > conventional Fenton (Pliego et al., 2015). Bio-electro-Fenton
has also been studied and utilized (Feng et al., 2010; Wang & Wang, 2017; Wang
et al., 2014; Xu, Zeng, Li, Zhang, & Sun, 2015), but in configurations that electro-
chemical processes are followed by the bio-treatments, the former may aim at an
increased bio-degradability of the effluent rather than complete mineralization of
contaminants (Mousset et al., 2016). Additional, the utilization of heterogeneous
22 H. HE AND Z. ZHOU

iron catalysts or complexing agents is promising to solve the issue of the low solu-
bility of ferric ion at high pH conditions (Duesterberg et al., 2008). Nanoparticulate
ZVI (Niu et al., 2012; Zhang et al., 2014), goethite (Lu, 2000), pyrite (Ammar et al.,
2015; Barhoumi et al., 2016; Bouzayani et al., 2017; Labiadh, Oturan, Panizza,
Hamadi, & Ammar, 2015), nanoparticulate Fe3O4 (Jiang, Sun, Feng, & Wang,
2016)) have been utilized as the heterogeneous iron catalysts. Heterogeneous reac-
tions can be more efficient because they consume less H2O2 per mole contaminant
destroyed, but in the meantime, they are much slower than corresponding aqueous
reactions (Pignatello et al., 2006). Further studies about heterogeneous electro-
Fenton are necessary. More investigations are in need with respect to catalyst car-
riers such as ligand (Keenan & Sedlak, 2008), activated carbon (Zhang et al., 2014),
Fe-loaded ion-exchange resin (Ba~ nuelos et al., 2013), reactive life time and adverse
consequences of nanomaterials, and techniques to prevents iron leaching.
Downloaded by [University of Florida] at 00:51 19 December 2017

Additional work to improve the cost-effectiveness of electro-Fenton processes


are needed. Development on reactor configuration may improve the overall perfor-
mance of electro-Fenton system, as a previous study indicated that flow through
configurations with three-dimensional electrodes may reduce the power input and
cost (Radjenovic & Sedlak, 2015). Reactor operation can be improved as well. Pre-
treatment with anodic oxidation and adsorption may reduce the COD concentra-
tion so electro-Fenton can be efficiently employed. Finally, a high capacity electro-
Fenton system is desired for scaled-up industrial applications. Dilution is usually
needed in existing electro-Fenton systems due to the limited capacity, but dilution
creates additional wastewater in large volumes and results in extra pumping costs.
Innovations on improving treatment efficiency with high concentrations of organic
pollutants will significantly improve the overall cost-effectiveness of electro-Fenton
processes.

Acknowledgment
The study was financially supported by Purdue Research Foundation Summer Faculty Grant.

ORCID
Zhi Zhou http://orcid.org/0000-0003-1252-2626

References
Ammar, S., Oturan, M. A., Labiadh, L., Guersalli, A., Abdelhedi, R., Oturan, N., and Brillas, E.
(2015). Degradation of tyrosol by a novel electro-Fenton process using pyrite as heteroge-
neous source of iron catalyst. Water Res., 74, 77–87. https://doi.org/10.1016/j.
watres.2015.02.006
Annabi, C., Fourcade, F., Soutrel, I., Geneste, F., Floner, D., Bellakhal, N., and Amrane, A.
(2016). Degradation of enoxacin antibiotic by the electro-Fenton process: Optimization, bio-
degradability improvement and degradation mechanism. J. Environ. Manage., 165, 96–105.
https://doi.org/10.1016/j.jenvman.2015.09.018
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 23

Anotai, J., Lu, M. C., and Chewpreecha, P. (2006). Kinetics of aniline degradation by Fenton
and electro-Fenton processes. Water Res., 40, 1841–1847. https://doi.org/10.1016/j.
watres.2006.02.033
Atmaca, E. (2009). Treatment of landfill leachate by using electro-Fenton method. J. Hazard.
Mater., 163, 109–114. https://doi.org/10.1016/j.jhazmat.2008.06.067
Babu, B. R., Venkatesan, P., Kanimozhi, R., and Basha, C. A. (2009). Removal of pharmaceuti-
cals from wastewater by electrochemical oxidation using cylindrical flow reactor and optimi-
zation of treatment conditions. J. Environ. Sci. Health A Tox. Hazard Subst. Environ. Eng.,
44, 985–994. https://doi.org/10.1080/10934520902996880
Babuponnusami, A., and Muthukumar, K. (2014). A review on Fenton and improvements to the
Fenton process for wastewater treatment. J. Environ. Chem. Eng., 2, 557–572. https://doi.org/
10.1016/j.jece.2013.10.011
Badawy, M. I., and Ali, M. E. (2006). Fenton’s peroxidation and coagulation processes for the
treatment of combined industrial and domestic wastewater. J. Hazard. Mater, 136, 961–966.
https://doi.org/10.1016/j.jhazmat.2006.01.042
Downloaded by [University of Florida] at 00:51 19 December 2017

Balci, B., Oturan, M. A., Oturan, N., and Sires, I. (2009). Decontamination of aqueous glypho-
sate, (aminomethyl)phosphonic acid, and glufosinate solutions by electro-Fenton-like pro-
cess with Mn2C as the catalyst. J. Agric. Food Chem., 57, 4888–4894. https://doi.org/10.1021/
jf900876x
Barb, W. G., Baxendale, J. H., George, P., and Hargrave, K. R. (1951). Reactions of ferrous and
ferric ions with hydrogen peroxide. Part I.—The ferrous ion reaction. Trans. Faraday Soc.,
47, 462–500. https://doi.org/10.1039/TF9514700462
Barhoumi, N., Oturan, N., Olvera-Vargas, H., Brillas, E., Gadri, A., Ammar, S., Oturan, M. A.
(2016). Pyrite as a sustainable catalyst in electro-Fenton process for improving oxidation of
sulfamethazine. Kinetics, mechanism and toxicity assessment. Water Res., 94, 52–61. https://
doi.org/10.1016/j.watres.2016.02.042
nuelos, J. A., Rodrıguez, F. J., Manrıquez, J. R., Bustos, E. A., Rodrıguez, J. C., Cruz, L. G.,
Ba~
Arriaga and L. A. Godınez (2013). Novel electro-fenton approach for regeneration of acti-
vated carbon. Environ Sci Technol, 47(14), 7927–7933.
Benitez, F. J., Acero, J. L., Real, F. J., Rubio, F. J., and Leal, A. I. (2001). The role of hydroxyl
radicals for the decomposition of p-hydroxy phenylacetic acid in aqueous solutions. Water
Res., 35, 1338–1343. https://doi.org/10.1016/S0043-1354(00)00364-X
Borras, N., Arias, C., Oliver, R., and Brillas, E. (2011). Mineralization of desmetryne by electro-
chemical advanced oxidation processes using a boron-doped diamond anode and an oxy-
gen-diffusion cathode. Chemosphere, 85, 1167–1175. https://doi.org/10.1016/j.chemosphere.
2011.09.008
Bossmann, S. H., Oliveros, E., G€ob, S., Siegwart, S., Dahlen, E. P., Payawan, L., Straub, M.,
Braun, A. M. (1998). New evidence against hydroxyl radicals as reactive intermediates in the
thermal and photochemically enhanced fenton reactions. J. Phys. Chem. A., 102, 5542–5550.
https://doi.org/10.1021/jp980129j
Bouzayani, B., Meijide, J., Pazos, M., Elaoud, S. C., and Sanroman, M. A. (2017). Removal of
polyvinylamine sulfonate anthrapyridone dye by application of heterogeneous electro-Fen-
ton process. Environ. Sci. Pollut. Res. Int., https://doi.org/10.1007/s11356-017-9468-5
Bray, W. C., and Gorin, M. H. (1932). Ferryl ion, a compound of tetravalent iron. J. Am. Chem.
Soc., 54, 2124–2125. https://doi.org/10.1021/ja01344a505
Brezonik, P., and Arnold, W. (2011). Water chemistry: An introduction to the chemistry of natu-
ral and engineered aquatic systems. New York: Oxford University Press.
Brillas, E., Banos, M. A., Skoumal, M., Cabot, P. L., Garrido, J. A., and Rodriguez, R. M. (2007).
Degradation of the herbicide 2,4-DP by anodic oxidation, electro-Fenton and photoelectro-
24 H. HE AND Z. ZHOU

Fenton using platinum and boron-doped diamond anodes. Chemosphere, 68, 199–209.
https://doi.org/10.1016/j.chemosphere.2007.01.038
Brillas, E., Boye, B., Sires, I., Garrido, J. A., Rodrıguez, R. M. A., Arias, C., Comninellis, C.
(2004). Electrochemical destruction of chlorophenoxy herbicides by anodic oxidation and
electro-Fenton using a boron-doped diamond electrode. Electrochim Acta, 49, 4487–4496.
https://doi.org/10.1016/j.electacta.2004.05.006
Brillas, E., and Casado, J. (2002). Aniline degradation by electro-Fenton and peroxi-coagulation
processes using a flow reactor for wastewater treatment. Chemosphere, 47, 241–248. https://
doi.org/10.1016/S0045-6535(01)00221-1
Brillas, E., Garcia-Segura, S., Skoumal, M., and Arias, C. (2010). Electrochemical incinera-
tion of diclofenac in neutral aqueous medium by anodic oxidation using Pt and boron-
doped diamond anodes. Chemosphere, 79, 605–612. https://doi.org/10.1016/j.
chemosphere.2010.03.004
Buxton, G., Greenstock, C., Helman, W., and Ross, A. (1988). Critical review of rate constants
for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (OH/O¡ in
Downloaded by [University of Florida] at 00:51 19 December 2017

aqueous solution. J. Phys. Chem. Ref. Data, 17, 513–886.


B€
uy€uks€ onmez, F., Rynk, R., Hess, T. F., and Bechinski, E. (1999). Occurrence, degradation and
fate of pesticides during composting. Compost Sci. Utili., 7, 66–82. https://doi.org/10.1080/
1065657X.1999.10701986
Ca~nizares, P., Garcıa-Gomez, J., Saez, C., and Rodrigo, M. (2004). Electrochemical oxidation of
several chlorophenols on diamond electrodes: Part II. Influence of waste characteristics and
operating conditions. J. Appl. Electrochem., 34, 87–94. https://doi.org/10.1023/B:
JACH.0000005587.52946.66
Casado, J., Fornaguera, J., and Galan, M. I. (2005). Mineralization of aromatics in water by sun-
light-assisted electro-fenton technology in a pilot reactor. Environ. Sci. Technol., 39, 1843–
1847. https://doi.org/10.1021/es0498787
Celebi, M., Oturan, N., Zazou, H., Hamdani, M., and Oturan, M. (2015). Electrochemical oxida-
tion of carbaryl on platinum and boron-doped diamond anodes using electro-Fenton tech-
nology. Sep. Purif. Technol., 156, 996–1002. https://doi.org/10.1016/j.seppur.2015.07.025
Choi, J. Y., Lee, Y. J., Shin, J., and Yang, J. W. (2010). Anodic oxidation of 1,4-dioxane on boron-
doped diamond electrodes for wastewater treatment. J. Hazard Mater., 179, 762–768. https://
doi.org/10.1016/j.jhazmat.2010.03.067
Chou, S., Huang, Y., Lee, S., Huang, G., and Huang, C. (1999). Treatment of high strength hex-
amine-containing wastewater by electro-Fenton method. Water Res., 33, 751–759. https://
doi.org/10.1016/S0043-1354(98)00276-0
Comninellis, C., and Nerini, A. (1995). Anodic oxidation of phenol in the presence of NaCl for
wastewater treatment. J. Appl. Electrochem., 25(1), 23–28.
Da Pozzo, A., Petrucci, E., and Merli, C. (2008). Electrogeneration of hydrogen peroxide in sea-
water and application to disinfection. J. Appl. Electrochem., 38, 997–1003. https://doi.org/
10.1007/s10800-008-9524-4
Daneshvar, N., Aber, S., Vatanpour, V., and Rasoulifard, M. H. (2008). Electro-Fenton treat-
ment of dye solution containing Orange II: Influence of operational parameters. J. Electroa-
nal. Chem., 615, 165–174. https://doi.org/10.1016/j.jelechem.2007.12.005
Fernandez de Dios, M.A.,  Rosales, E., Fernandez-Fernandez, M., Pazos, M., and Sanroman,
M. (2015). Degradation of organic pollutants by heterogeneous electro-Fenton process
using Mn-alginate composite. J. Chem. Technol. Biotechnol., 90, 1439–1447. https://doi.
org/10.1002/jctb.4446
De Laat, J., Truong Le, G., and Legube, B. (2004). A comparative study of the effects of chloride,
sulfate and nitrate ions on the rates of decomposition of H2O2 and organic compounds by
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 25

Fe(II)/H2O2 and Fe(III)/H2O2. Chemosphere, 55, 715–723. https://doi.org/10.1016/j.


chemosphere.2003.11.021
Deguillaume, L., Leriche, M., and Chaumerliac, N. (2005). Impact of radical versus non-radical
pathway in the Fenton chemistry on the iron redox cycle in clouds. Chemosphere, 60, 718–
724. https://doi.org/10.1016/j.chemosphere.2005.03.052
Deng, Y., and Englehardt, J. D. (2006). Treatment of landfill leachate by the Fenton process.
Water Res., 40, 3683–3694. https://doi.org/10.1016/j.watres.2006.08.009
Ding, X., Ai, Z., and Zhang, L. (2012). Design of a visible light driven photo-electrochemical/
electro-Fenton coupling oxidation system for wastewater treatment. J. Hazard. Mater., 239–
240, 233–240. https://doi.org/10.1016/j.jhazmat.2012.08.070
Do, J. S., and Chen, C. P. (1994). In situ oxidative degradation of formaldehyde with hydrogen
peroxide electrogenerated on the modified graphites. J. Appl. Electrochem., 24, 936–942.
https://doi.org/10.1007/BF00348785
Duesterberg, C. K., Mylon, S. E., and Waite, T. D. (2008). pH effects on iron-catalyzed oxidation
using Fenton’s reagent. Environ. Sci. Technol., 42, 8522–8527. https://doi.org/10.1021/
Downloaded by [University of Florida] at 00:51 19 December 2017

es801720d
Duesterberg, C. K., and Waite, T. D. (2006). Process optimization of fenton oxidation using
kinetic modeling. Environ. Sci. Technol., 40, 4189–4195. https://doi.org/10.1021/es060311v
Edelahi, M., Oturan, N., Oturan, M., Padellec, Y., Bermond, A., and El Kacemi, K. (2003). Deg-
radation of diuron by the electro-Fenton process. Environ. Chem. Lett., 1, 233–236. https://
doi.org/10.1007/s10311-003-0052-5
El-Ghenymy, A., Cabot, P. L., Centellas, F., Garrido, J. A., Rodrıguez, R. M., Arias, C., Brillas, E.
(2013). Mineralization of sulfanilamide by electro-Fenton and solar photoelectro-Fenton in
a pre-pilot plant with a Pt/air-diffusion cell. Chemosphere, 91, 1324–1331. https://doi.org/
10.1016/j.chemosphere.2013.03.005
Faust, B., and Hoigne, J. (1990). Photolysis of Fe(III) hydroxy complexes as sources of OH radi-
cals in clouds, fog, and rain. Atmos. Environ. Part A-Gen. Top., 24, 79–89. https://doi.org/
10.1016/0960-1686(90)90443-Q
Feng, C. H., Li, F. B., Mai, H. J., and Li, X. Z. (2010). Bio-electro-Fenton process driven by
microbial fuel cell for wastewater treatment. Environ. Sci. Technol., 44, 1875–1880. https://
doi.org/10.1021/es9032925
Feng, L., Oturan, N., van Hullebusch, E. D., Esposito, G., and Oturan, M. A. (2014). Degradation
of anti-inflammatory drug ketoprofen by electro-oxidation: Comparison of electro-Fenton
and anodic oxidation processes. Environ. Sci. Pollut. Res. Int., 21, 8406–8416. https://doi.org/
10.1007/s11356-014-2774-2
Fernandes, A., Morao, A., Magrinho, M., Lopes, A., and Goncalves, I. (2004). Electrochemical
degradation of C. I. Acid orange 7. Dyes Pigments, 61, 287–296. https://doi.org/10.1016/j.
dyepig.2003.11.008
Gao, G., Zhang, Q., Hao, Z., and Vecitis, C. D. (2015). Carbon nanotube membrane stack for
flow-through sequential regenerative electro-Fenton. Environ. Sci. Technol., 49, 2375–2383.
https://doi.org/10.1021/es505679e
Garcia, O., Isarain-Chavez, E., Garcia-Segura, S., Brillas, E., and Peralta-Hernandez, J. (2013).
Degradation of 2,4-dichlorophenoxyacetic acid by electro-oxidation and electro-Fenton/
BDD processes using a pre-pilot plant. Electrocatalysis, 4, 224–234. https://doi.org/10.1007/
s12678-013-0135-4
Ghoneim, M., El-Desoky, H., and Zidan, N. (2011). Electro-Fenton oxidation of Sunset Yellow
FCF azo-dye in aqueous solutions. Desalination, 274, 22–30. https://doi.org/10.1016/j.
desal.2011.01.062
26 H. HE AND Z. ZHOU

Gogate, P., and Pandit, A. (2004). A review of imperative technologies for wastewater treatment
I: Oxidation technologies at ambient conditions. Adv. Environ. Res., 8, 501–551. https://doi.
org/10.1016/S1093-0191(03)00032-7
Gozzo, F. (2001). Radical and non-radical chemistry of the Fenton-like systems in the presence
of organic substrates. J. Mol. Catal. A-Chem., 171, 1–22. https://doi.org/10.1016/S1381-1169
(01)00099-1
Gumus, D., and Akbal, F. (2016). Comparison of Fenton and electro-Fenton processes for oxi-
dation of phenol. Process Saf. Environ. Prot., 103, 252–258. https://doi.org/10.1016/j.
psep.2016.07.008
Haag, W., and Yao, C. (1992). Rate constants for reaction of hydroxyl radicals with several drinking
water contaminants. Environ. Sci. Technol., 26, 1005–1013. https://doi.org/10.1021/es00029a021
Hammami, S., Oturan, N., Bellakhal, N., Dachraoui, M., and Oturan, M. (2007). Oxidative deg-
radation of direct orange 61 by electro-Fenton process using a carbon felt electrode: Applica-
tion of the experimental design methodology. J. Electroanal. Chem., 610, 75–84. https://doi.
org/10.1016/j.jelechem.2007.07.004
Downloaded by [University of Florida] at 00:51 19 December 2017

Hou, B., Han, H., Jia, S., Zhuang, H., Xu, P., and Li, K. (2016). Three-dimensional heteroge-
neous electro-Fenton oxidation of biologically pretreated coal gasification wastewater using
sludge derived carbon as catalytic particle electrodes and catalyst. J. Taiwan Inst. Chem.
Eng., 60, 352–360. https://doi.org/10.1016/j.jtice.2015.10.032
Huang, Y. H., Huang, Y. F., Chang, P. S., and Chen, C. Y. (2008). Comparative study of oxida-
tion of dye-Reactive Black B by different advanced oxidation processes: Fenton, electro-Fen-
ton and photo-Fenton. J. Hazard. Mater., 154, 655–662. https://doi.org/10.1016/j.
jhazmat.2007.10.077
Hug, S. J., and Leupin, O. (2003). Iron-catalyzed oxidation of arsenic(III) by oxygen and by
hydrogen peroxide: PH-dependent formation of oxidants in the Fenton reaction. Environ.
Sci. Technol., 37, 2734–2742. https://doi.org/10.1021/es026208x
Ito, K., Jian, W., Nishijima, W., Baes, A., Shoto, E., and Okada, M. (1998). Comparison of ozon-
ation and AOPs combined with biodegradation for removal of THM precursors in treated
sewage effluents. Water Sci. Technol., 38, 179–186.
Jasmann, J. R., Borch, T., Sale, T. C., and Blotevogel, J. (2016). Advanced electrochemical oxida-
tion of 1,4-dioxane via dark catalysis by novel titanium dioxide (TiO2) pellets. Environ. Sci.
Technol., 50, 8817–8826. https://doi.org/10.1021/acs.est.6b02183
Jiang, H., Sun, Y., Feng, J., and Wang, J. (2016). Heterogeneous electro-Fenton oxidation of azo
dye methyl orange catalyzed by magnetic Fe3O4 nanoparticles. Water Sci. Technol., 74,
1116–1126. https://doi.org/10.2166/wst.2016.300
Jones, C. W., Braithwaite, M. J., and Clark, J. H. (1999). Applications of hydrogen peroxide and
derivatives. Cambridge: Royal Society of Chemistry.
Kang, Y., Cho, M., and Hwang, K. (1999). Correction of hydrogen peroxide interference on
standard chemical oxygen demand test. Water Res., 33, 1247–1251. https://doi.org/10.1016/
S0043-1354(98)00315-7
Kavitha, V., and Palanivelu, K. (2005). Destruction of cresols by Fenton oxidation process.
Water Res., 39, 3062–3072. https://doi.org/10.1016/j.watres.2005.05.011
Kayan, B., G€ ozmen, B., Demirel, M., and Gizir, A. M. (2010). Degradation of acid red 97 dye in
aqueous medium using wet oxidation and electro-Fenton techniques. J. Hazard Mater, 177,
95–102. https://doi.org/10.1016/j.jhazmat.2009.11.076
Keenan, C. R., and Sedlak, D. L. (2008). Ligand-enhanced reactive oxidant generation by nano-
particulate zero-valent iron and oxygen. Environ. Sci. Technol., 42, 6936–6941. https://doi.
org/10.1021/es801438f
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 27

Kishimoto, N., Kitamura, T., Kato, M., and Otsu, H. (2013). Reusability of iron sludge as an iron
source for the electrochemical Fenton-type process using Fe2C/HOCl system. Water Res., 47,
1919–1927. https://doi.org/10.1016/j.watres.2013.01.021
Kremer, M. (1999). Mechanism of the Fenton reaction. Evidence for a new intermediate. Phys.
Chem. Chem. Phys., 1, 3595–3605. https://doi.org/10.1039/a903915e
Kwan, W. P., and Voelker, B. M. (2002). Decomposition of hydrogen peroxide and organic
compounds in the presence of dissolved iron and ferrihydrite. Environ. Sci. Technol., 36,
1467–1476. https://doi.org/10.1021/es011109p
Kwon, B., Lee, D., Kang, N., and Yoon, J. (1999). Characteristics of p-chlorophenol oxida-
tion by Fenton’s reagent. Water Res., 33, 2110–2118. https://doi.org/10.1016/S0043-
1354(98)00428-X
Labiadh, L., Oturan, M. A., Panizza, M., Hamadi, N. B., and Ammar, S. (2015). Complete
removal of AHPS synthetic dye from water using new electro-Fenton oxidation catalyzed by
natural pyrite as heterogeneous catalyst. J. Hazard. Mater., 297, 34–41. https://doi.org/
10.1016/j.jhazmat.2015.04.062
Downloaded by [University of Florida] at 00:51 19 December 2017

Lee, C., Lee, E., Lim, Y., Park, K., Park, H., and Lim, D. (2017). Enhanced electrochemical oxida-
tion of phenol by boron-doped diamond nanowire electrode. RSC Adv., 7, 6229–6235.
https://doi.org/10.1039/C6RA26287B
Lee, E., Lee, H., Kim, Y., Sohn, K., and Lee, K. (2011). Hydrogen peroxide interference in chemi-
cal oxygen demand during ozone based advanced oxidation of anaerobically digested live-
stock wastewater. Int. J. Environ. Sci. Technol., 8, 381–388. https://doi.org/10.1007/
BF03326225
Li, X., Zhang, P., Jin, L., Shao, T., Li, Z., and Cao, J. (2012). Efficient photocatalytic decomposi-
tion of perfluorooctanoic acid by indium oxide and its mechanism. Environ. Sci. Technol.,
46, 5528–5534. https://doi.org/10.1021/es204279u
Liao, C. H., Kang, S. F., and Wu, F. A. (2001). Hydroxyl radical scavenging role of chloride and
bicarbonate ions in the H2O2/UV process. Chemosphere, 44, 1193–1200. https://doi.org/
10.1016/S0045-6535(00)00278-2
Lin, H., Niu, J., Xu, J., Huang, H., Li, D., Yue, Z., and Feng, C. (2013). Highly efficient and mild
electrochemical mineralization of long-chain perfluorocarboxylic acids (C9-C10) by Ti/
SnO2-Sb-Ce, Ti/SnO2-Sb/Ce-PbO2, and Ti/BDD electrodes. Environ. Sci. Technol., 47,
13039–13046. https://doi.org/10.1021/es4034414
Lin, S., Lin, C., and Leu, H. (1999). Operating characteristics and kinetic studies of surfactant
wastewater treatment by Fenton oxidation. Water Res., 33, 1735–1741. https://doi.org/
10.1016/S0043-1354(98)00403-5
Lin, S., and Lo, C. (1997). Fenton process for treatment of desizing wastewater. Water Res., 31,
2050–2056. https://doi.org/10.1016/S0043-1354(97)00024-9
Lu, M. C. (2000). Oxidation of chlorophenols with hydrogen peroxide in the presence of goe-
thite. Chemosphere, 40(2), 125–130.
Liu, H., Wang, C., Li, X., Xuan, X., Jiang, C., and Cui, H. (2007). A novel electro-Fenton process
for water treatment: Reaction-controlled pH adjustment and performance assessment. Envi-
ron. Sci. Technol., 41, 2937–2942. https://doi.org/10.1021/es0622195
Liu, X., Yang, D., Zhou, Y., Zhang, J., Luo, L., Meng, S., Tang, L. (2017). Electrocatalytic proper-
ties of N-doped graphite felt in electro-Fenton process and degradation mechanism of levo-
floxacin. Chemosphere, 182, 306–315. https://doi.org/10.1016/j.chemosphere.2017.05.035
Liu, Y., Chen, S., Quan, X., Yu, H., Zhao, H., and Zhang, Y. (2015). Efficient mineralization of
perfluorooctanoate by electro-Fenton with H2O2 electro-generated on hierarchically porous
carbon. Environ. Sci. Technol., 49, 13528–13533. https://doi.org/10.1021/acs.est.5b03147
28 H. HE AND Z. ZHOU

Liu, Y., Xie, J., Ong, C. N., Vecitis, C., and Zhou, Z. (2015). Electrochemical wastewater treat-
ment with carbon nanotube filters coupled with in situ generated H2O2. Environ. Sci.-Water
Res. Technol., 1, 769–778. https://doi.org/10.1039/C5EW00128E
Ma, L., Zhou, M., Ren, G., Yang, W., and Liang, L. (2016). A highly energy-efficient flow-
through electro-Fenton process for organic pollutants degradation. Electrochim. Acta, 200,
222–230. https://doi.org/10.1016/j.electacta.2016.03.181
Martinez-Huitle, C. A., and Ferro, S. (2006). Electrochemical oxidation of organic pollutants for
the wastewater treatment: Direct and indirect processes. Chem. Soc. Rev., 35, 1324–1340.
https://doi.org/10.1039/B517632H
Mollah, M. Y., Schennach, R., Parga, J. R., and Cocke, D. L. (2001). Electrocoagulation (EC)–sci-
ence and applications. J. Hazard Mater., 84, 29–41. https://doi.org/10.1016/S0304-3894(01)
00176-5
Moon, D., Ezuka, M., Maruyama, T., Osakada, K., and Yamamoto, T. (1993). Kinetic study on
chemical oxidation of leucoemeraldine base polyaniline to emeraldine base. Macromolecules,
26, 364–369. https://doi.org/10.1021/ma00054a016
Downloaded by [University of Florida] at 00:51 19 December 2017

Mousset, E., Wang, Z., and Lefebvre, O. (2016). Electro-Fenton for control and removal of
micropollutants – process optimization and energy efficiency. Water Sci. Technol., 74, 2068–
2074. https://doi.org/10.2166/wst.2016.353
Nidheesh, P., and Gandhimathi, R. (2012). Trends in electro-Fenton process for water and
wastewater treatment: An overview. Desalination, 299, 1–15. https://doi.org/10.1016/j.
desal.2012.05.011
Nidheesh, P., Gandhimathi, R., Velmathi, S., and Sanjini, N. (2014). Magnetite as a heteroge-
neous electro Fenton catalyst for the removal of Rhodamine B from aqueous solution. RSC.
Adv., 4, 5698–5708. https://doi.org/10.1039/c3ra46969g
Niu, J., Lin, H., Xu, J., Wu, H., and Li, Y. (2012). Electrochemical mineralization of perfluorocar-
boxylic acids (PFCAs) by ce-doped modified porous nanocrystalline PbO2 film electrode.
Environ. Sci. Technol., 46, 10191–10198.
Oturan, N., Panizza, M., and Oturan, M. A. (2009). Cold incineration of chlorophenols in aque-
ous solution by advanced electrochemical process electro-Fenton. Effect of number and posi-
tion of chlorine atoms on the degradation kinetics. J. Phys. Chem. A, 113, 10988–10993.
https://doi.org/10.1021/jp9069674
Ozcan, A., Sahin, Y., Koparal, A. S., and Oturan, M. A. (2008). Degradation of picloram by the
electro-Fenton process. J. Hazard. Mater., 153, 718–727. https://doi.org/10.1016/j.
jhazmat.2007.09.015

Ozcan, A., Şahin, Y., Savaş Koparal, A., and Oturan, M. A. (2008). Carbon sponge as a new
cathode material for the electro-Fenton process: Comparison with carbon felt cathode and
application to degradation of synthetic dye basic blue 3 in aqueous medium. J. Electroanal.
Chem., 616, 71–78. https://doi.org/10.1016/j.jelechem.2008.01.002
Panizza, M., and Cerisola, G. (2008). Electrochemical degradation of methyl red using BDD and
PbO2 anodes. Ind. Eng. Chem. Res., 47, 6816–6820. https://doi.org/10.1021/ie8001292
Panizza, M., and Cerisola, G. (2009). Electro-Fenton degradation of synthetic dyes. Water Res.,
43, 339–344. https://doi.org/10.1016/j.watres.2008.10.028
Pignatello, J. (1992). Dark and photo assisted Fe3C catalyzed degradation of chlorophenoxy her-
bicides by hydrogen peroxide. Environ. Sci. Technol., 26, 944–951. https://doi.org/10.1021/
es00029a012
Pignatello, J., Oliveros, E., and Mackay, A. (2006). Advanced oxidation processes for organic
contaminant destruction based on the Fenton reaction and related chemistry. Crit. Rev. Envi-
ron. Sci. Technol., 36, 1–84. https://doi.org/10.1080/10643380500326564
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 29

Pimentel, M., Oturan, N., Dezotti, M., and Oturan, M. (2008). Phenol degradation by advanced
electrochemical oxidation process electro-Fenton using a carbon felt cathode. Appl. Catal. B-
Environ., 83, 140–149. https://doi.org/10.1016/j.apcatb.2008.02.011
Pliego, G., Zazo, J., Garcia-Munoz, P., Munoz, M., Casas, J., and Rodriguez, J. (2015). Trends in
the intensification of the Fenton process for wastewater treatment: An overview. Crit. Rev.
Environ. Sci. Technol., 45, 2611–2692. https://doi.org/10.1080/10643389.2015.1025646
Pratap, K., and Lemley, A. T. (1998). Fenton electrochemical treatment of aqueous atrazine and
metolachlor. J. Agric. Food Chem., 46, 3285–3291. https://doi.org/10.1021/jf9710342
Primo, O., Rivero, M., and Ortiz, I. (2008). Photo-Fenton process as an efficient alternative to
the treatment of landfill leachates. J. Hazard. Mater., 153, 834–842. https://doi.org/10.1016/j.
jhazmat.2007.09.053
Qiang, Z., Chang, J. H., and Huang, C. P. (2002). Electrochemical generation of hydrogen per-
oxide from dissolved oxygen in acidic solutions. Water Res., 36, 85–94. https://doi.org/
10.1016/S0043-1354(01)00235-4
Qiang, Z., Chang, J. H., and Huang, C. P. (2003). Electrochemical regeneration of Fe2C in Fen-
Downloaded by [University of Florida] at 00:51 19 December 2017

ton oxidation processes. Water Res., 37, 1308–1319. https://doi.org/10.1016/S0043-1354(02)


00461-X
Qiu, S., He, D., Ma, J., Liu, T., and Waite, T. (2015). Kinetic modeling of the electro-fenton pro-
cess: Quantification of reactive oxygen species generation. Electrochim. Acta, 176, 51–58.
https://doi.org/10.1016/j.electacta.2015.06.103
Qu, Y., Zhang, C., Li, F., Chen, J., and Zhou, Q. (2010). Photo-reductive defluorination of per-
fluorooctanoic acid in water. Water Res., 44, 2939–2947. https://doi.org/10.1016/j.
watres.2010.02.019
Radjenovic, J., and Sedlak, D. L. (2015). Challenges and opportunities for electrochemical pro-
cesses as Next-generation technologies for the treatment of contaminated water. Environ.
Sci. Technol., 49, 11292–11302. https://doi.org/10.1021/acs.est.5b02414
Ratanatamskul, C., Masomboon, N., and Lu, M. (2011). Application of fenton processes for deg-
radation of aniline. Pestic. Mod. World – Pestic. Manag., https://doi.org/10.5772/17323
Ribeiro, A. R., Nunes, O. C., Pereira, M. F., and Silva, A. M. (2015). An overview on the
advanced oxidation processes applied for the treatment of water pollutants defined in the
recently launched Directive 2013/39/EU. Environ. Int., 75, 33–51. https://doi.org/10.1016/j.
envint.2014.10.027
Rivas, F. J., Beltran, F. J., Frades, J., and Buxeda, P. (2001). Oxidation of p-hydroxybenzoic acid by
Fenton’s reagent. Water Res., 35, 387–396. https://doi.org/10.1016/S0043-1354(00)00285-2
Rivas, F. J., Beltran, F., Gimeno, O., and Carvalho, F. (2003). Fenton-like oxidation of landfill
leachate. J. Environ. Sci. Health A Tox Hazard Subst. Environ. Eng., 38, 371–379. https://doi.
org/10.1081/ESE-120016901
Rosales, E., Pazos, M., Longo, M. A., and Sanroman, M. A. (2009). Influence of operational
parameters on electro-Fenton degradation of organic pollutants from soil. J. Environ. Sci.
Health A Tox Hazard Subst. Environ. Eng., 44, 1104–1110. https://doi.org/10.1080/
10934520903005111
Rosales, E., Pazos, M., and Sanroman, M. A. (2012). Advances in the Electro-fenton process for
remediation of recalcitrant organic compounds. Chem. Eng. Technol., 35, 609–617. https://
doi.org/10.1002/ceat.201100321
Ruiz, E. J., Arias, C., Brillas, E., Hernandez-Ramırez, A., and Peralta-Hernandez, J. M. (2011).
Mineralization of Acid Yellow 36 azo dye by electro-Fenton and solar photoelectro-Fenton
processes with a boron-doped diamond anode. Chemosphere, 82, 495–501. https://doi.org/
10.1016/j.chemosphere.2010.11.013
30 H. HE AND Z. ZHOU

Ruiz, F., Martinez, P., Castro, E., Humana, R., Peretti, H., and Visintin, A. (2013). Effect of elec-
trolyte concentration on the electrochemical properties of an AB5-type alloy for Ni/MH bat-
teries. Int. J. Hydrogen Energy, 38, 240–245. https://doi.org/10.1016/j.ijhydene.2012.10.007
Sedlak, D., and Andren, A. (1991). Oxidation of chlorobenzene with Fenton reagent. Environ.
Sci. Technol., 25, 777–782. https://doi.org/10.1021/es00016a024
Shan, Z., Lu, M., Wang, L., MacDonald, B., MacInnis, J., Mkandawire, M., Oakes, K. D. (2016).
Chloride accelerated Fenton chemistry for the ultrasensitive and selective colorimetric detec-
tion of copper. Chem. Commun. (Camb), 52, 2087–2090. https://doi.org/10.1039/
C5CC07446K
Shin, Y. U., Yoo, H. Y., Kim, S., Chung, K. M., Park, Y. G., Hwang, K. H., Lee, J. (2017). Sequen-
tial combination of electro-Fenton and electrochemical chlorination processes for treatment
of anaerobically-digested food wastewater. Environ. Sci. Technol., 51(18), 10700–10710.
https://doi.org/10.1021/acs.est.7b02018
Sires, I., Brillas, E., Oturan, M. A., Rodrigo, M. A., and Panizza, M. (2014). Electrochemical
advanced oxidation processes: Today and tomorrow. A review. Environ. Sci. Pollut. Res. Int.,
Downloaded by [University of Florida] at 00:51 19 December 2017

21, 8336–8367. https://doi.org/10.1007/s11356-014-2783-1


Sires, I., Oturan, N., Oturan, M., Rodriguez, R., Garrido, J., and Brillas, E. (2007). Electro-Fenton
degradation of antimicrobials triclosan and triclocarban. Electrochim. Acta, 52, 5493–5503.
https://doi.org/10.1016/j.electacta.2007.03.011
Skoumal, M., Rodriguez, R., Cabot, P., Centellas, F., Garrido, J., Arias, C., Brillas, E. (2009). Elec-
tro-Fenton, UVA photoelectro-Fenton and solar photoelectro-Fenton degradation of the
drug ibuprofen in acid aqueous medium using platinum and boron-doped diamond anodes.
Electrochim. Acta, 54, 2077–2085. https://doi.org/10.1016/j.electacta.2008.07.014
Stumm, W., and Morgan, J. J. (1985). Aquatic chemistry: An introduction emphasizing chemical
equilibria in natural waters. New York, N.Y.: Wiley.
Sun, Y., and Pignatello, J. (1993). Photochemical reactions involved in the total mineralization of 2,4-
D by Fe3C/H2O2/UV. Environ. Sci. Technol., 27, 304–310. https://doi.org/10.1021/es00039a010
Tian, J., Olajuyin, A. M., Mu, T., Yang, M., and Xing, J. (2016). Efficient degradation of rhoda-
mine B using modified graphite felt gas diffusion electrode by electro-Fenton process. Envi-
ron. Sci. Pollut. Res. Int., 23, 11574–11583. https://doi.org/10.1007/s11356-016-6360-7
Ting, W. P., Lu, M. C., and Huang, Y. H. (2008). The reactor design and comparison of Fenton,
electro-Fenton and photoelectro-Fenton processes for mineralization of benzene sulfonic
acid (BSA). J. Hazard. Mater., 156, 421–427. https://doi.org/10.1016/j.jhazmat.2007.12.031
Ting, W. P., Lu, M. C., and Huang, Y. H. (2009). Kinetics of 2,6-dimethylaniline degrada-
tion by electro-Fenton process. J. Hazard Mater., 161, 1484–1490. https://doi.org/
10.1016/j.jhazmat.2008.04.119
Tokumura, M., Sugawara, A., Raknuzzaman, M., Habibullah-Al-Mamun, M., and Masunaga, S.
(2016). Comprehensive study on effects of water matrices on removal of pharmaceuticals by
three different kinds of advanced oxidation processes. Chemosphere, 159, 317–325. https://
doi.org/10.1016/j.chemosphere.2016.06.019
Umar, M., Aziz, H. A., and Yusoff, M. S. (2010). Trends in the use of Fenton, electro-Fenton and
photo-Fenton for the treatment of landfill leachate. Waste Manag., 30, 2113–2121. https://
doi.org/10.1016/j.wasman.2010.07.003
Verma, S., Khandegar, V., and Saroha, A. (2013). Removal of chromium from electroplating
industry effluent using electrocoagulation. J. Hazard., Toxic, Radioactive Waste, 17, 146–
152. https://doi.org/10.1061/(ASCE)HZ.2153-5515.0000170
Vermilyea, A., and Voelker, B. (2009). Photo-fenton reaction at near neutral pH. Environ. Sci.
Technol., 43, 6927–6933. https://doi.org/10.1021/es900721x
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 31

Wang, B. B., Cao, M. H., Tan, Z. J., Wang, L. L., Yuan, S. H., and Chen, J. (2010). Photochemical
decomposition of perfluorodecanoic acid in aqueous solution with VUV light irradiation. J.
Hazard. Mater., 181, 187–192. https://doi.org/10.1016/j.jhazmat.2010.04.115
Wang, C. T., Hu, J. L., Chou, W. L., and Kuo, Y. M. (2008). Removal of color from real dyeing
wastewater by Electro-Fenton technology using a three-dimensional graphite cathode. J.
Hazard Mater., 152, 601–606. https://doi.org/10.1016/j.jhazmat.2007.07.023
Wang, C., Chou, W., Chung, M., and Kuo, Y. (2010). COD removal from real dyeing wastewater
by electro-Fenton technology using an activated carbon fiber cathode. Desalination, 253,
129–134. https://doi.org/10.1016/j.desal.2009.11.020
Wang, N., Zheng, T., Zhang, G., and Wang, P. (2016). A review on Fenton-like processes for
organic wastewater treatment. J. Environ. Chem. Eng., 4, 762–787. https://doi.org/10.1016/j.
jece.2015.12.016
Wang, X. Q., Liu, C. P., Yuan, Y., and Li, F. B. (2014). Arsenite oxidation and removal driven by
a bio-electro-Fenton process under neutral pH conditions. J. Hazard Mater., 275, 200–209.
https://doi.org/10.1016/j.jhazmat.2014.05.003
Downloaded by [University of Florida] at 00:51 19 December 2017

Wang, X., Chen, S., Gu, X., and Wang, K. (2009). Pilot study on the advanced treatment of land-
fill leachate using a combined coagulation, fenton oxidation and biological aerated filter pro-
cess. Waste Manag., 29, 1354–1358. https://doi.org/10.1016/j.wasman.2008.10.006
Wang, Y. T., and Wang, R. S. (2017). A Bio-electro-fenton system employing the composite
FePc/CNT/SS316 Cathode. Materials (Basel), 10(2), E169.
Wang, Y., Liu, Y., Liu, T., Song, S., Gui, X., Liu, H., and Tsiakaras, P. (2014). Dimethyl phthalate
degradation at novel and efficient electro-Fenton cathode. Appl. Catal. B-Environ., 156, 1–7.
https://doi.org/10.1016/j.apcatb.2014.02.041
Wang, Y., Zhao, H., Chai, S., Wang, Y., Zhao, G., and Li, D. (2013). Electrosorption enhanced
electro-Fenton process for efficient mineralization of imidacloprid based on mixed-valence
iron oxide composite cathode at neutral pH. Chem. Eng. J., 223, 524–535. https://doi.org/
10.1016/j.cej.2013.03.016
Xu, N., Zeng, Y., Li, J., Zhang, Y., and Sun, W. (2015). Removal of 17 beta-estrodial in a bio-
electro-Fenton system: Contribution of oxidation and generation of hydroxyl radicals with
the Fenton reaction and carbon felt cathode. RSC. Adv., 5, 56832–56840. https://doi.org/
10.1039/C5RA08053C
Zazo, J. A., Casas, J. A., Mohedano, A. F., Gilarranz, M. A., and Rodriguez, J. J. (2005). Chemical
pathway and kinetics of phenol oxidation by Fenton’s reagent. Environ. Sci. Technol., 39,
9295–9302. https://doi.org/10.1021/es050452h
Zhang, C., Zhou, L., Yang, J., Yu, X., Jiang, Y., and Zhou, M. (2014). Nanoscale zero-valent iron/
AC as heterogeneous Fenton catalysts in three-dimensional electrode system. Environ. Sci.
Pollut. Res. Int., 21, 8398–8405. https://doi.org/10.1007/s11356-014-2791-1
Zhang, H., Choi, H. J., and Huang, C. P. (2005). Optimization of Fenton process for the treat-
ment of landfill leachate. J. Hazard Mater., 125, 166–174. https://doi.org/10.1016/j.
jhazmat.2005.05.025
Zhang, H., Choi, H. J., and Huang, C. P. (2006). Treatment of landfill leachate by Fenton’s
reagent in a continuous stirred tank reactor. J. Hazard. Mater., 136, 618–623. https://doi.org/
10.1016/j.jhazmat.2005.12.040
Zhang, H., Fei, C., Zhang, D., and Tang, F. (2007). Degradation of 4-nitrophenol in aqueous
medium by electro-Fenton method. J. Hazard. Mater., 145, 227–232. https://doi.org/
10.1016/j.jhazmat.2006.11.016
Zhang, H., Zhang, D., and Zhou, J. (2006). Removal of COD from landfill leachate by electro-
Fenton method. J. Hazard. Mater., 135, 106–111. https://doi.org/10.1016/j.
jhazmat.2005.11.025
32 H. HE AND Z. ZHOU

Zhou, M., Yu, Q., Lei, L., and Barton, G. (2007). Electro-Fenton method for the removal of
methyl red in an efficient electrochemical system. Sep. Purif. Technol., 57, 380–387. https://
doi.org/10.1016/j.seppur.2007.04.021
Zhuo, Q., Deng, S., Yang, B., Huang, J., and Yu, G. (2011). Efficient electrochemical oxidation of
perfluorooctanoate using a Ti/SnO2-Sb-Bi anode. Environ. Sci. Technol., 45, 2973–2979.
https://doi.org/10.1021/es1024542
Zuo, Y., and Hoigne, J. (1992). Formation of hydrogen peroxide and depletion of oxalic acid in
atmospheric water by photolysis of iron(III) oxalate complexes. Environ. Sci. Technol., 26,
1014–1022. https://doi.org/10.1021/es00029a022
Downloaded by [University of Florida] at 00:51 19 December 2017

You might also like