You are on page 1of 35

Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

Contents lists available at ScienceDirect

Journal of Photochemistry and Photobiology C:


Photochemistry Reviews
journal homepage: www.elsevier.com/locate/jphotochemrev

Applied photoelectrocatalysis on the degradation of organic


pollutants in wastewaters
Sergi Garcia-Segura ∗ , Enric Brillas ∗
Laboratori d’Electroquímica dels Materials i del Medi Ambient, Departament de Química Física, Facultat de Química, Universitat de Barcelona, Martí i
Franquès 1-11, 08028 Barcelona, Spain

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

Article history: A large variety of electrochemical advanced oxidation processes (EAOPs) have been recently developed
Received 20 October 2016 to remove organic pollutants from wastewaters to avoid their serious health-risk factors from their envi-
Received in revised form 16 January 2017 ronmental accumulation and to reuse the treated water for human activities. The effectiveness of EAOPs
Accepted 30 January 2017
is based on the in situ production of strong reactive oxygen species (ROS) like hydroxyl radical (• OH).
Available online 9 February 2017
Photoelectrocatalysis (PEC) has emerged as a promising powerful EAOP by combining photocatalytic and
electrolytic processes. It consists in the promotion of electrons from the valence band to the conduction
Keywords:
band of a semiconductor photocatalyst upon light irradiation, with production of positive holes. The fast
Hydroxyl radical
Organics degradation
recombination of the electron/hole pairs formed is avoided in PEC by applying an external bias potential
Photocatalysis to the photocatalyst that extracts the photogenerated electrons up to the cathode of the electrolytic cell.
Photoelectrocatalysis Organics can be oxidized directly by the holes, • OH formed from water oxidation with holes and other
TiO2 photoanode ROS produced between the electrons and dissolved O2 . This paper presents a general and critical review
Wastewater treatment on the application of PEC to the remediation of wastewaters with organic pollutants. Special attention
is made over the different kinds of photocatalysts utilized and preparation methods of the most ubiqui-
tous TiO2 materials. Typical PEC systems and main operation variables that affect the effectiveness of the
degradation process are also examined. An exhaustive analysis of the advances obtained on the treat-
ment of dyes, chemicals and pharmaceuticals from synthetic solutions, as well as of real wastewaters, is
performed. Finally, research prospects are proposed for the future development of PEC with perspectives
to industrial application.
© 2017 Elsevier B.V. All rights reserved.

Abbreviations: A, absorbance of the most intense UV/Vis peak; A0 , initial absorbance of the most intense UV/Vis peak; ACF, activated carbon fiber; Ag/AgCl, reference
electrode of Ag/AgCl (KCl saturated); ALD, atomic layer deposition; AOP, advanced oxidation process; APS, atmospheric plasma spray; BDD, boron-doped diamond; BOD5 ,
biochemical oxygen demand at 5 days; c, organic concentration (mM or mg L−1 ); c0 , initial organic concentration (mM or mg L−1 ); CB, conductive band; CVD, chemical vapor
deposition; COD, chemical oxygen demand (mg O2 L−1 ); COD0 , initial chemical oxygen demand (mg O2 L−1 ); DO, direct ozonation; DP, direct photolysis; DRS, diffuse
reflectance spectroscopy; DSA , dimensionally stable anode; EAOP, electrochemical advanced oxidation process; e− CB , electron in the conduction band; EDS, energy dispersive
®

spectrometry; EF, electro-Fenton; EIS, electrochemical impedance spectroscopy; EO, electrochemical oxidation; Eanod , anodic potential (V); Ebg , band gap energy (eV); Ecat ,
cathodic potential (V); Ecell , potential difference of the cell (V); Efb , flat band energy (V); FESEM, field-emission SEM; FTO, fluor-doped tin dioxide; h+ VB , hole in the valence
band; HPLC, high-performance liquid chromatography; I, current (A or mA); ITO, indium-tin oxide; janod , anodic current density (mA cm−2 ); ␭, wavelength of UV/Vis spectrum;
␭max , maximum wavelength of UV/Vis spectrum; LC–MS, liquid chromatography–mass spectrometry; NB, nanobelt; NT, nanotube; NTA, nanotube array; PANI, polyaniline; PC,
photocatalysis; PEC, photoelectrocatalysis; PEF, photoelectro-Fenton; PTFE, polytetrafluoroethylene; PZC, point of zero charge; ROS, reactive oxygen species; SCE, saturated
calomel electrode; SCL, space charge layer; SEM, scanning electron microscopy; SPEC, solar photoelectrocatalysis; t, electrolysis time; TEM, transmission electron microscopy;
TOC, total organic carbon (mg C L−1 ); TOC0 , initial total organic carbon (mg C L−1 ); VB, valence band; UV, ultraviolet; UVA, ultraviolet A; UVB, ultraviolet B; UVC, ultraviolet
C; Vis, visible; XRD, X-ray diffraction; XPS, X-ray photoelectron spectroscopy.
∗ Corresponding authors.
E-mail addresses: sergigarcia@ub.edu (S. Garcia-Segura), brillas@ub.edu (E. Brillas).

http://dx.doi.org/10.1016/j.jphotochemrev.2017.01.005
1389-5567/© 2017 Elsevier B.V. All rights reserved.
2 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Fundamentals of photoelectrocatalysis (PEC) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3. Experimental conditions for PEC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.1. Photocatalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.1.1. TiO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.1.2. WO3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.1.3. ZnO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.1.4. Other semiconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.1.5. Modified TiO2 materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.2. Preparation of TiO2 photocatalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2.1. Thin-film electrodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2.2. Nanostructured materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.2.3. Characterization of synthesized photocatalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.3. PEC systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.4. Operation parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
4. Destruction of organic pollutants by PEC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.1. Dyes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4.1.1. TiO2 photoanodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4.1.2. Doped TiO2 and composites with TiO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.1.3. Other photocatalytic materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.2. Chemicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.2.1. TiO2 and composites with TiO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2.2. Other photoanodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4.3. Pharmaceuticals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.4. Real wastewaters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5. Conclusions and prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
Acknowlegments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

Dr. Sergi Garcia-Segura is a researcher dedicated to


the development of Electrochemical Advanced Oxidation
Processes to remove organic pollutants from environ- Dyes, chemicals and pharmaceuticals are some of the most com-
ment such photoelectrocatalysis. He has conducted his mon recalcitrant organic pollutants. For instance, many industries
research at the Universitat de Barcelona (Spain), Univer-
sity of Queensland (Australia), Universidade Federal do
including textile, cosmetic, paper, leather, light-harvesting, arrays,
Rio Grande do Norte (Brazil), Chia Nan University of Phar- agricultural research, photoelectrochemical cells, pharmaceutical
macy and Science (Taiwan), Bonn Universität (Germany) and food produce large volumes of wastewater polluted with high
and Arizona State University (USA). He received the
concentration of dyes and other components. As a result, about
International Society of Electrochemistry Prize for Envi-
ronmental Electrochemistry 2014, the Green Talents 280,000 tons of textile dyes are currently discharged in effluents
award 2015 and the Antonio Aldaz prize 2016. He has every year and introduced in the aquatic environment [5]. This
published 40 peer-reviewed papers (h-index = 19). has induced many governments to apply legislation that prescribes
and limits the emission of pollutants. To face this environmen-
Dr. Enric Brillas is Full Professor of Physical Chem- tal problem, three different approaches have been considered: (i)
istry at the Universitat de Barcelona since 1987. He was the development of Green Chemical and Technological Processes,
president of the Electrochemistry Group of the Real
Sociedad Española de Química (2004–2008). His research
(ii) the use of the 3R (reduce, reuse and recycle) sustainability
is pre-eminently devoted on organic electrochemistry, consciousness and (iii) the application of wastewater remedia-
photocatalysis, photoelectrocatalysis, electrocatalysis and tion technologies. The latter approach has received great attention
electrochemical treatments of organic pollutants. He
received the Chemviron Carbon 1995 and the CIDETEC
because it is easily usable and can solve the contamination prob-
2014 Awards. Associate Editor of Chemosphere and mem- lems.
ber of the Editorial Board of Journal of Hazardous Materials Current methods for wastewater treatment have been based on
and Applied Catalysis B: Environmental. He has published
oxidation processes including physicochemical, biological, chem-
330 peer-reviewed papers (h-index = 61) and 22 books
and book chapters, and presented 280 communications ical and electrochemical treatments [5]. Note that no universal
to scientific congresses. strategy on wastewater remediation is feasible because of the
extremely diverse composition of industrial waste that usually con-
tains a complex mixture of organic and inorganic compounds and
1. Introduction mainly depends on the nature and concentration of pollutants.
Physicochemical techniques require high cost of equipment and
One of the main current worldwide concerns is the growth of usually present low effectiveness, particularly over dyes and phar-
water pollution by organic compounds arising from many indus- maceuticals. Biological treatments are environmentally friendly,
trial, agricultural and urban human activities. The vast majority of produce less sludge than physicochemical systems and are rela-
these compounds are persistent organic pollutants, owing to their tively inexpensive. Nevertheless, their application is rather limited
resistance to conventional chemical, biological and photolytic pro- since treatment needs a large land area, has sensitivity toward tox-
cesses. As a result, they have been detected in rivers, lakes, oceans icity of certain chemicals and operation time is very long. Over
and even drinking water all over the world. This constitutes a seri- the past three decades, many advanced oxidation processes (AOPs)
ous environmental health problem mainly due to their toxicity and have been developed as more effective technologies to remove per-
potential hazardous health effects (carcinogenicity, mutagenicity sistent organic pollutants from wastewaters [6]. AOPs are based on
and bactericidality) on living organisms, including human beings the in situ production of highly reactive hydroxyl radicals (• OH)
[1–4]. that non-selectively react with most organics and are able to
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 3

degrade even highly recalcitrant compounds [5–7]. This radical a 140


is the second strongest oxidant known after fluorine, displaying
a high standard reduction potential of E◦ (• OH/H2 O) = 2.80 V (SHE)

Number of publications
120
and rate constants for reaction with several contaminants in the
order of 106 to 1010 M−1 s−1 [6,8]. • OH has so short lifetime, a 100
few nanoseconds in water [9], that can be rapidly self-eliminated 80
from the treatment system. The most ubiquiotous AOPs are chem-
ical, photochemical and photocatalytic systems such as H2 O2 with 60
UVC radiation (H2 O2 /UVC), ozone and ozone based processes (O3 , 40
O3 /UVC, O3 /H2 O2 and O3 /H2 O2 /UVC), titanium dioxide based pro-
cesses (TiO2 /UV and TiO2 /H2 O2 /UV) and Fenton’s reaction based
20
methods (Fenton (Fe2+ /H2 O2 ) and photo-Fenton (Fe2+ /H2 O2 /UV)) 0
[10]. Photocatalysis (PC) uses a semiconductor, usually TiO2 known 2001- 2005- 2008- 2011- 2014-
as the photocatalyst, under light illumination (UV or solar) to 2004 2007 2010 2013 2016
generate electron/hole pairs with ability to degrade most organic b 200
pollutants by producing the strong oxidant • OH at its surface.

Number of publications
This technique has prominent advantages including non-toxicity,
160
low cost, no secondary pollution and thorough mineralization.
However, PC is restricted by its low photonic efficiency. The fast
120
recombination of photogenerated electron/hole pairs at the photo-
catalyst surface represents the major drawback for PC applications.
80
Over the last two decades, electrochemical advanced oxidation
processes (EAOPs) have gained increasing attention as a promising
class of AOPs [5,6,11,12]. They have emerged as novel treatment 40
technologies for the elimination of a broad-range of organic con-
taminants from wastewaters. Several advantages of EAOPs include 0
TiO Modified WO ZnO Other
high energy efficiency, amenability to automation, simple equip- 2
TiO
3
photoanodes
ment, safety because they operate under mild conditions (ambient 2

temperature and pressure) and versatility. From the beginning of c 200


the XXI Century, the combination of electrochemistry and PC in the
Number of publications

so-called photoelectrocatalysis (PEC) method has deserved increas- 160


ing attention. It is based on a semiconductor photoanode that is
irradiated by light with energy equal or greater than its band gap 120
and simultaneously biased by a gradient of potential. PEC offers the
opportunity to separate the charges of the photogenerated elec- 80
tron/hole pairs, strongly enhancing the mineralization of organic
pollutants in wastewaters. Note that urban and industrial wastew- 40
aters usually present conductivity enough to effectively perform
the PEC treatment of their organic pollutants because these water 0
matrices already contain electrolytes like salts of sulfate, chloride Dyes Chemicals Drugs Real
and carbonate. Since PEC is still in development, most research wastewaters
has been made using synthetic wastewater prepared with ultra-
pure water and addition of electrolytes. If electrolytes are added Fig. 1. Number of photoelectrocatalysis (PEC) publications over treatment of
organic pollutants in wastewaters as a function of: (a) published year, (b) photoan-
for real PEC application, they should be removed from the final
ode used and (c) degraded pollutant.
treated solution before disposal. PEC has also been successfully
applied to inorganic ion reduction, microorganism inactivation,
CO2 reduction, and production of electricity and hydrogen from firstly exposed, followed by an overview about its experimen-
water splitting [13–15]. When sunlight is used as energy source, it tal conditions including the photocatalysts used, the preparation
is known as solar PEC (SPEC). methods for the most common TiO2 materials, the kinds of PEC
Fig. 1a shows the increasing number of papers over PEC classi- systems utilized and the main operation parameters that affect the
fied for wastewater remediation since 2001, showing the growing degradation performance. The application of PEC to the treatment
interest over its development. Fig. 1b highlights that TiO2 is the of dyes, chemicals, pharmaceuticals and real wastewaters, remark-
preferred material as photoanode in PEC, although modified TiO2 ing the use of different photoanodes, is discussed. Examples are
and to smaller extent WO3 , ZnO and other materials have been uti- given for the different kinds of organics to evidence the advances
lized as photoanodes as well. Special attention has been made on achieved in this technique.
the applications with TiO2 and its modified materials because they Several reviews on the environmental application of PEC have
offer important advantages as a result of their non-toxic, optical, been published based on the advances of this technology up
low-cost and biocompatibility properties. Thus, PEC has been used to 2011–2012 [16–18]. The present paper shows a much more
to remove a large variety of organic pollutants, most of them dyes detailed review from the research mainly devoted on the latter five
and to less extent chemicals, drugs and real wastewaters, as can years, where PEC has achieved more mature developments in the
be seen in Fig. 1c. The large development of PEC over the last 15 following points: (i) the preparation of new photoanodes not only
years has open novel ways for the application of this EAOP to water based on TiO2 , but also on WO3 , ZnO and other photocatalytic mate-
remediation. rials involving ␤-PbO2 , BiVO4 , BiPO4 and ␣-Fe2 O3 , among others,
The aim of this paper is to present a general and critical review (ii) the design of new PEC systems for using direct illumination with
on the use of the PEC technology for the remediation of organic sunlight and filter-press flow cells at lab-scale, (iii) the coupling
pollutants in wastewaters. The fundamentals of the technique are of PEC with O3 and other EAOPs to enhance its oxidation power,
4 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

and (iv) the use of PEC to treat pharmaceuticals and real wastew- improved if it is deposited onto a conductive substrate to act as
aters, pre-eminently reported from 2012. Moreover, this review a photoanode in a photoelectrolytic system leading to the PEC pro-
describes comprehensively the main characteristics or organics cess [18,20–24].
degradation by PEC such as their kinetic decay, mineralization rate The PEC process on water splitting was firstly described by
and degree, identification of aromatic and cyclic intermediates, as Fujishima and Honda on 1972 and analyzed from the electrochem-
well as final carboxylic acid generated, and reaction sequences pro- ical point of view by Brockis et al. [33,34] on the early 1980s, but
posed. The significant effect of operation parameters over organic it was not up to the beginning of XXI Century when it was applied
degradation is extensively documented and exemplified in selected to wastewater treatment, as stated above. It uses an electrolytic
figures for better understanding. system containing a thin-film active photoanode subjected to light
illumination with application of a constant bias potential to the
2. Fundamentals of photoelectrocatalysis (PEC) anode (Eanod ), a constant cell potential (Ecell ) or a constant anodic
current density (janod ). This promotes the extraction of photoin-
The PEC technique combines both electrolytic and photocat- duced e− CB by the external electrical circuit, thereby yielding an
alytic processes and has received considerable attention because of efficient separation of the e− CB /h+ VB pairs since Reactions (3)–(7)
its ability to retard the recombination of electron-hole (e− CB /h+ VB ) are inhibited [18,20–24]. The prevention of charge recombination
pairs, increasing the lifetime of the holes. The basic process of PEC upgrades the photocatalytic efficiency of the anode with genera-
consists in the ejection of an electron (e− CB ) from the valence band tion of higher quantities of holes by reaction (1) and acceleration of
(VB) of a semiconductor, which is fully filled, to the conductive band organics oxidation compared to classical PC. The lifetime of holes
(CB), which is completely empty, generating a positively charged is increased and they have more opportunities either to directly
vacancy or hole (h+ VB ). The band gap is related to the light irradia- oxidize the organic pollutants adsorbed on the photoanode surface
tion utilized. The semiconductor has to be exposed to an irradiation or indirectly destroy them with more amounts of • OH formed from
with greater energy than that of its band gap (Ebg ), giving rise to the Reaction (2). In PEC, the photocatalyst can be easily recovered after
photoexcitation of the e− CB from VB to CB. The light then allows the usage and recycled for consecutive treatments.
generation of e− CB /h+ VB pairs via Reaction (1) [5,18–24]: The material more extensively used as semiconductor photocat-
alyst in PEC is the anatase crystalline form of TiO2 , since it can act as
Semiconductor + hv → e− CB + h+ VB (1)
an active photoanode with low cost, low toxicity, high stability and
h+
Photogenerated VB is a strong oxidizing species, whereas the wide band gap of 3.2 eV. Fig. 2 illustrates the mechanism of the pro-
promoted e− CB is a potential reductor. Organic pollutants are then cesses taking place in an n-type semiconductor such as TiO2 [18].
oxidized by the photogenerated h+ VB up to their complete min- The e− CB /h+ VB pairs are produced by irradiating photons of UV light
eralization. It is also proposed the reaction of h+ VB with adsorbed with h␯ > Ebg of TiO2 . The photogenerated electrons flow through
water to form the strong oxidant • OH from Reaction (2) that min- the external circuit due to a potential gradient from the photoan-
eralizes the organic pollutants, although there is no clear evidence ode to the cathode. The holes can thus attack directly to organics R
for the formation of free hydroxyl radical from h+ VB . The e− CB can and/or generate high amounts of • OH via photo-oxidation of water
react with adsorbed O2 to form the superoxide radical O2 •− from in much larger extent than ROS formed from e− CB via O2 photore-
Reaction (3). duction because most electrons are lost from the photoanode to the
cathode [18].
h+ VB + H2 O → • OH + H+ (2) As shown in Fig. 2, a typical cathode involves the reduction of
− •− water to H2 . However, it is feasible to enhance the oxidation power
e CB + O2 → O2 (3)
of PEC by using carbonaceous cathodes that allow the in situ elec-
Other weaker reactive oxygen species (ROS) such as H2 O2 and trogeneration of H2 O2 from two-electron reduction of injected O2
hydroperoxyl radical HO2 • can be produced via Reactions (4) and by Reaction (8) [5,11]. The presence of H2 O2 in the electrolytic
(5): medium upgrades the production of oxidant • OH from its reduction
O2 •− + H+ → HO2 • (4) with the promoted e− CB by Reaction (9) [34–38]. Although Reaction
(9) is accepted for authors working in PC and PEC, a recent research

2HO2 → H2 O2 + O2 (5) has reported controversial results over its validity [39].
Nevertheless, the promoted e−
CB is an unstable species of an O2 + 2H+ + 2 e− → H2 O2 (8)
excited state and tends to return to the ground state either with
adsorbed • OH by Reaction (6) or pre-eminently by recombination e− CB + H2 O2 → • OH + OH− (9)
with the unreacted h+ VB from Reaction (7)) [18–24].
Apart from the above photolytic processes, a semiconductor M
e− CB + • OH → OH− (6)
submitted to an anodic potential can remove the organics by elec-
e− CB + h+ VB → Catalyst + heat (7) trochemical oxidation (EO). This is feasible because it oxidizes the
water to adsorbed hydroxyl radical M(• OH) as follows [40–42]:
The last reaction represents the main drawback for the efficient
use of absorbed photons in the classical PC. To improve the abate- M + H2 O → M(• OH) + H+ + e− (10)
ment of organic pollutants from wastewaters by this technique,
the separation of charges formed from Reaction (1) has been per- Unfortunately, the low conductivity of semiconductors used in
formed using nanoparticulated photocatalysts with high specific PEC only allows the pass of small janod values, usually < 10 mA cm−2 ,
area in suspension in the effluent, but, unfortunately, the recovery leading to low M(• OH) production with very poor ability for the
of these materials after treatment is complicated [25–27]. Research mineralization of organic pollutants. In the PEC treatment of a
efforts to solve this problem have then been devoted to the synthe- wastewater, it is very difficult to know the contribution of EO to the
sis of nanoparticulated photocatalysts immobilized onto different overall process, since the photogeneration of electron/hole pairs
substrates [28–31]. The fixation of photocatalysts on supports pro- from Reaction (1) competes with M(• OH) generation from Reaction
duces an inevitable and significant reduction of their active specific (10) at the photocatalyst surface, although the large predominance
area with the consequent drop in pollutant removal. However, of the former process explains the much greater oxidation ability of
the efficiency of the immobilized photocatalyst can be strongly PEC than EO [31]. The use of high janod values in PEC causes normally
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 5

Fig. 2. Mechanism for PEC process using TiO2 photocatalyst. Adapted with permission from ref. [18].

a loss of its performance as a result of a change of the photocatalyst photogenerated electrons are dragged out by an external electrical
properties, as will be discussed below. circuit (see Fig. 2) [18].

3. Experimental conditions for PEC 3.1.2. WO3


WO3 is other n-type semiconductor metal oxide as TiO2 . The
Most papers devoted to the PEC treatment of organic pollutants crystalline structure of WO3 is temperature dependent, although
in wastewaters present an experimental study in consecutive steps. the most common is the monoclinic phase between 17 and 330 ◦ C
Firstly, the pollutant and photocatalyst are selected. Since the pho- [48]. The Ebg values of tungsten oxides are around 2.5–2.7 eV, which
toactive material is not commercially available, its synthesis and are appreciably lower than those of TiO2 , and for this reason, WO3
characterization are subsequently described. Finally, the PEC sys- presents better performance under visible light irradiation due to
tem used to explore the process performance is detailed and the the easier formation of photogenerated electron/hole pairs [49,50].
degradation results are reported as function of the operation vari- The main drawback of this material is that it is not as innocuous as
ables chosen. This section gives an overview all these experimental TiO2 , since it is a hazardous, toxic and irritant metal oxide [51]. The
conditions, which are essential to understand the potential appli- first usage of WO3 photoanode for wastewater remediation was
cation of PEC. reported by Hepel and Luo in 2001 [52,53], who well-proven the
potential applicability of this material to the abatement of azo dyes.
This possibility has been recently explored over different pollutants
3.1. Photocatalysts by a moderate number of articles, as can be seen in Fig. 1b.

The PEC efficiency on organic pollutant remediation is directly


related to the semiconductors selected as photoanodes by their 3.1.3. ZnO
intrinsic photocatalytical properties. In general, they are metal ZnO is extensively found in the nature and is considerably
oxides with an appropriate Ebg value between the fulfilled VB cheaper than the above photocatalysts. It is also considered as an
and the empty CB to photogenerate electron/hole pairs upon light environment-friendly material because of its innocuous character
irradiation. The different materials used in PEC for wastewater over the health of living beings [54]. ZnO presents two crys-
treatment are detailed below. talline structures, the hexagonal wurtzite and the cubic zinc blende.
Wurtzite is the most thermodynamically stable structure at ambi-
ent conditions [55]. Its Ebg value is 3.4 eV, with high electrochemical
3.1.1. TiO2 stability and possible use under natural sunlight irradiation [56,57].
TiO2 is the most utilized metal oxide photocatalyst for envi- ZnO thin films are transparent and this improves the light pen-
ronmental remediation applications, as shown in Fig. 1b. It is etration into the material and hence, the photogeneration of
an n-type semiconductor with three main crystalline structures, electron/hole pairs from Reaction (1). Although the number of pub-
namely rutile, anatase and brookite. The former is the most ther- lications related to ZnO as photoanode in PEC is rather limited (see
modynamically stable phase, whereas the metastable anatase and Fig. 1b), Li et al. [58] have reported that it can present better perfor-
brookite phases can be transformed into the most stable rutile mance even than TiO2 , thus appearing as an alternative functional
one under annealing in the range 600−800◦ C [43,44]. The Ebg for material for PEC application to wastewater treatment.
TiO2 is slightly superior to 3.0 eV, with little differences between
its crystalline phases, with values of 3.02 eV for rutile, 3.23 eV for
anatase and 3.14 eV for brookite [45,46]. However, anatase with 3.1.4. Other semiconductors
the higher Ebg value is the most active phase in PC upon UV irra- Other many pure metal oxide semiconductors have been
diation [45,47]. This has been justified by the prolonged lifetime of checked as photoanode materials in PEC aiming to search inex-
the photogenerated charge carriers (e− CB and h+ VB ) and the spatial pensive, efficient and highly stable photocatalysts to the electrical
charge separation promoted by anatase crystalline structure [45]. current with ability to absorb greater energy from solar spectrum
This behavior is very different under PEC conditions where the per- to photoinduce electron/hole pairs. Nevertheless, the performance
formance between the different phases is quite similar since the of such materials cannot be easily compared with that found with
6 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

classical TiO2 because different PEC systems and operation vari-


ables are used for environmental remediation.
The hematite ␣-Fe2 O3 has been considered one of these alter-
native candidates because of its low cost, chemical stability and
high light absorbance. Its low Ebg = 2.2 eV allows its direct applica-
tion to visible light [59,60]. Other promising photocatalyst is MnO
with a very low Ebg = 1.3 eV, being an interesting functional material
with low cost, large surface, electrochemical stability and innocu-
ous character [61,62]. The combination of MnO with polyaniline
demonstrated an enhancement of its PEC properties owing to the
interaction with the polyaniline bandgap of 2.8 eV that reduces the
recombination reaction of electrons and holes [62]. On the other
hand, SnO2 is a chemically and thermally stable semiconductor
with so wide Ebg of 3.5–3.8 eV that is difficultly used as photo-
catalyst [63]. Pure non-conductive SnO2 can be easily doped with
Sb improving considerably its electrical conductivity, although the
resulting Sb-SnO2 material employed as photocatalyst presents low
electrochemical stability with short lifetime upon current appli-
cation [64]. Other semiconductor such as ␤-PbO2 has also been
utilized as photocatalyst due to its very small Ebg = 1.4 eV [65], but
this material is only useful in alkaline medium since under acidic
conditions, it leaches toxic lead ions to the medium. In contrast, sta-
ble bismuth materials like BiVO4 with Ebg = 2.5 eV [66] and Bi2 WO6
with Ebg = 2.8 eV [67] under visible light irradiation and BiPO4 with
Ebg = 3.8 eV [68] under UV illumination have shown an excellent
effectiveness and good performance for the PEC treatment of some
organic pollutants.

3.1.5. Modified TiO2 materials


The PEC performance of a photocatalyst mainly depends on
the following factors: (i) the light absorption properties, (ii) the
reduction and oxidation rates on the surface by the photogener-
ated electrons and holes, and (iii) the recombination rate of such
charges. To enhance the PEC activity of the most ubiquiotous TiO2
photocatalyst, various strategies have been developed including: (i)
the construction of TiO2 nanotubes (NTs), nanotube arrays (NTAs),
nanobelts (NBs) or nanorods structures [69,70], (ii) the doping with
metals like Cr [71], Cu [72] and Fe [73] or non-metals like B [74]
and N [75] (second generation of photocalysts), (iii) the synthesis
of composites with metals like Pd [76], Ag [77] and Au [78], other
metal oxides like SiO2 [79], WO3 [80], Fe3 O4 [81], SnO2 [82] and
Cu2 O [83], metal sulfides like CdS [84] and Sb2 S3 [85], and car-
bonaceous materials like carbon cloth and graphene [86], and (iv) Fig. 3. SEM micrographs for thin films of: (a) TiO2 thin film deposited onto stainless
steel by atmospheric plasma spray. (Reproduced with permission from Ref. [31])
new titanium compounds like TiNbO5 [87] and the cubic double-
and (b) TiO2 nanotube arrays (NTAs) prepared by anodization of a Ti foil in 10 wt%
perovskite CaCu3 Ti4 O12 [88] (third generation of photocalysts). water + 0.5 wt% NH4 F in glycerol at 30 V for 50 h (Reproduced with permission from
TiO2 thin films prepared on substrates like Ti, stainless steel, Ref. [89]).
indium-tin oxide (ITO) and fluor-doped tin dioxide (FTO), among
others, by several procedures have been extensively used in PEC.
These films have granulate, compact and low porosity structure,
as exemplified in the scanning electron microscopy (SEM) micro- the same preparation technique, Liao et al. [69] found an Ebg = 2.7 eV
graph of Fig. 3 [32,89], and this restricts the light absorption to UV for crystalline anatase Ti/TiO2 NTAs, a value much lower than 3.2 eV
radiation with high recombination rate of the photogenerated elec- of a granulate anatase thin film, and that allowed an enhanced PEC
tron/hole pairs, as stated above. Since 2007, highly crystalline and activity using visible light provided by a 60 W incandescent lamp.
ordered TiO2 nanostructures with large surface area and porosity The doping of TiO2 with metals or non-metals introduces new
in the form of nanotubes, nanobelts or nanorods have been synthe- VB and/or CB related to the impurity energy levels that enhance the
sized to improve the PEC activity of thin-film photoanodes. A higher degradation of organic pollutants by PEC compared with undoped
efficiency for organics removal is described using thin nanostruc- photocatalyst. Kerkez et al. [72] synthesized TiO2 nanorod arrays
tures as a result of a synergistic effect of absorbing visible light, films deposited onto FTO doped with CuO up to 0.26% molar of Cu2+
larger semiconductor/electrolyte interface due to higher surface (Cu2+ -TiO2 ). The Ebg value of these materials decreased gradually
area, minority carriers generated within a distance from the surface with raising dopant proportion, from 3.1 eV for the bare TiO2 /FTO
equal to the sum of the width of the depletion layer and the diffu- to 2.6 eV for the higher doped material, which allowed operating
sion length escape recombination, and smaller recombination of under visible light improving PEC treatment of organics. This was
electron/hole pairs [89]. Fig. 3b shows a SEM micrograph of a typi- explained by the transference of the photogenerated e− CB in TiO2
cal nanotubular TiO2 photoanode synthesized by Ti anodization and to the less energetic CB of CuO, producing Cu+ from Reaction (11).
composed of one-dimensional crystalline NTAs with uniform diam- Cu2+ sites then act as traps of photogenerated e− CB increasing the
eters of about 100 nm and wall thicknesses near 10 nm [89]. Using electron/hole lifetime, and the resulting Cu+ can be oxidized to Cu2+
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 7

Fig. 4. Proposed PEC mechanism for: (a) doped B-TiO2 NTs (Reproduced with permission from ref. [74]) and (b) Sb2 S3 /TiO2 composite (Reproduced with permission from
Ref. [85]).

either at the photoanode or with O2 via Reaction (12) to form O2 •− distance between the photocatalyst surface and the beginning of
that can originate the sequence of ROS by Reactions (4)–(5). the flat bands. No SCL exists at the flat band potential (Efb ), where
all bands are flats, and its value grews with increasing the applied
Cu2+ + e− CB → Cu+ (11) Eanod respect to Efb .
+ Other strategy for improving the TiO2 photoactivity is the prepa-
Cu + O2 → Cu 2+
+ O2 •− (12)
ration of composites with noble metals and other semiconductors.
A different behavior was found when doping TiO2 with a non- An enhancement of the PEC process has been found by decorat-
metal compound. An interesting case is the doping of TiO2 NTs ing TiO2 NTs with Pd, Ag and Au nanoparticles [76–78]. This has
with B in the form of NaBF4 . For this photocatalyst, Bessegato et al. been ascribed to the lower work function of the noble metal than
[74] determined two Ebg values, one of 3.3 eV related to TiO2 and the electron affinity of TiO2 NTs that originates a Schottky bar-
another of 2.2 eV due to the formation of intermediary energy lev- rier potential in the interfaces making energetically favorable the
els between the VB and CB of TiO2 created by substituted B atoms. electron transfer from the CB of the semiconductor to the metal
The proposed PEC mechanism with absorption of UV (for TiO2 ) and and significantly reducing the recombination of electron/hole pairs
visible light (from B dopant level) is illustrated in Fig. 4a. This fig- [76]. However, the authors do not consider the feasible simultane-
ure also shows the bending of all energetic bands upon application ous EO of organics by physisorbed hydroxyl radicals formed at the
of a bias potential due to the change of the Fermi potential of the metal surface from water oxidation by Reaction (10). The integral
semiconductor [74,85]. This bending leads to the formation of a treatment by PEC and EO has been reported for a SnO2 /TiO2 com-
space charge layer (SCL) or depletion layer, characterized by the posite by assembling sieve-like macroporous Sb-doped SnO2 film
depletion of photogenerated electrons and corresponding to the on vertically aligned TiO2 NTs [82]. In contrast, the strong hetero-
8 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

ing the dipping of the foil into the colloidal suspension, drying and
annealing at 100–300 ◦ C for ca. 3 h. Although the resulting Ti/TiO2
photocatalysts are the easiest applicable ones at lab scale, the coat-
ings obtained by this method present low stability because they
are easily cracked by the gas evolution at the photoanode surface
during PEC treatment. Low janod values are needed to be applied
to upgrade their operational life. The alternative use of the sol-
gel coating laser calcination method allows an improvement of the
adherence of TiO2 coatings onto Ti [93].
Other interesting method is the spray painting and thermal
Fig. 5. Schematic diagram illustrating the transfer of the photogenerated holes and decomposition [94,95], which consists in spraying a solution of
electrons in a TiO2 /graphene/Cu2 O interface. titanyl acetylacetonate (C10 H14 O5 Ti) onto a conductive substrate.
Source: Reproduced with permission from ref. [86]. The solution concentration, solution flow rate, nozzle-to-substrate
distance, and substrate temperature are crucial parameters that
junction between TiO2 and other n-type mixed oxides causes larger affect the coating properties. Clear and colorless thin films are then
photoactivity of their bicomposites. Fig. 4b shows a diagram of the consistently formed through thermal decomposition of the pre-
PEC mechanism for Ti/TiO2 NTAs decorated with Sb2 S3 nanoparti- cursor under oxygen. The painted coating could also be further
cles in orthorhombic phase (stibinite), with Ebg values of 3.2 and annealed at 400- 600 ◦ C to favor the thermal decomposition and
1.7 eV, respectively [85]. The photogenerated electrons in Sb2 S3 by to oxidize the precursors giving a best oxide layer coating. This
UV and visible light are injected into the CB of TiO2 because of the methodology is faster than the sol-gel method, allowing coating
more cathodic potential of the former (-0.77 V) respect to the lat- larger substrate surfaces in appreciably lower times. However, the
ter (−0.50 V). Furthermore, the photogenerated holes in the VB of obtained Ti/TiO2 thin-film photoanodes still present low mechan-
TiO2 are transferred to that of Sb2 S3 , creating a high concentra- ical stability like in the case of sol-gel methodology.
tion of holes in the sensitizer/electrolyte interfaces. This decreases TiO2 thin films can also be prepared by magnetron sputtering,
the recombination of the electron/hole pair and increases the PEC a physical vapor deposition technology based on a plasma coating
efficiency process, along with the advantage that Sb2 S3 acts as a process [96,97]. The sputtering material or precursor is ejected due
sensitizer practically in the entire visible spectrum. The PEC behav- to bombardment of ions to the target surface in a vacuum cham-
ior of an interesting TiO2 /graphene/Cu2 O tricomposite has been ber. The ionized plasma beam is accurately directed and focused
recently explored by Yang et al. [86]. As can be seen in Fig. 5, both in front of the target of the substrate by magnetic fields, giving
TiO2 NTAs and Cu2 O (a p-type semiconductor with Ebg = 2.2 eV) can excellent layer uniformity and smooth coatings. One strong point
be excited by UV/Vis light to produce electron/holes pairs. The elec- of this technology is that any conducting substrate can be sputtered
trons of the CB of TiO2 then across the inserted graphene layer and without decomposition or coating mechanical failure, because no
recombine with the holes of the VB of Cu2 O. In this way, oxidants substrate heading is required. He et al. [97] prepared TiO2 coating
• OH can be formed from the holes of VB of TiO from Reaction (2)
2 over ITO and Ti substrates. The main disadvantages of magnetron
without possibility of recombination with the electrons of the CB sputtering are: (i) the slow deposition speed which makes the pro-
of Cu2 O, which can originate the chain of ROS from the starting cess more expensive, (ii) only small surfaces can be coated, and (iii)
Reaction (3) with O2 . the low adhesion of the coatings that could diminish the opera-
The use of new titanium-metal oxides like CaCu3 Ti4 O12 with tion life of photoanodes, although it is considerably superior to the
very low Ebg = 1.5 eV [88] and high stability represents a promising conventional sol-gel coatings.
way to be explored in the next future for improving the PEC process Chemical vapor deposition (CVD) is an extensively used coating
of organic pollutants in wastewaters. technology for a wide range of applications [98]. A precursor gas
is flowed into a chamber containing one or more heated objects
3.2. Preparation of TiO2 photocatalysts to be coated, occurring the chemical reactions to form the coat-
ing on/near the hot surface to be coated. The chemical by-products
One of the most crucial points of PEC for the remediation of required are exhausted out of the chamber along with the unre-
polluted waters is the preparation of stable and re-utilizable pho- acted precursor gases usually at sub-torr total pressures and at
tocatalysts onto conducting materials as support. Great research temperatures typically ranging from 200 to 1600 ◦ C [98]. CVD yields
efforts have been carried out to develop novel methodologies better adherence of the TiO2 coating onto the conductive sub-
to synthesize a large variety of photoanodes. This subsection is strate than sol-gel, and it allows controlling the film thickness from
devoted to detail the most important methodologies for the prepa- nanometers up to micrometers as well as the crystallinity degree of
ration of TiO2 thin-film and nanostructured photocatalysts, since the coating, although it is much more expensive. The main limita-
they are the most used materials, as stated above. Unfortunately, tions of CVD are related to the higher cost of the precursors and the
the possible effect of the supporting material on the photocatalytic kinds of substrates with smaller surface areas that can be coated.
response is barely addressed in the literature and will not be con- The different thermal expansion coefficients between the substrate
sidered here. and the coating can stress the deposited films causing mechani-
cal failure or coating breaking. The first application of CVD coated
3.2.1. Thin-film electrodes photoanodes in PEC remediation of wastewaters was reported by
The sol-gel method is the most common technique to synthesize Hitchman and Tian [99], who prepared TiO2 thin films by low pres-
TiO2 thin-film photocatalysts [90–92]. It consists in the preparation sure between 257 and 400 ◦ C using titanium tetra-tertbutoxide as
of a titanium dioxide colloidal suspension by mixing titanium(IV) chemical precursor. The coatings obtained presented great photo-
isopropoxide (Ti(i-OPr)4 ) with glacial acetic acid under constant catalytic activity and stability, allowing their reuse.
stirring and keeping a molar ratio H+ /Ti = 4, followed by dilution in Atomic layer deposition (ALD) is other chemical thin-film tech-
2-propanol with 1:1 Ti/alcohol ratio. A stable colloidal suspension nology in which the substrate surface is exposed to gaseous
solution is then obtained by adding water and HNO3 under stirring precursors [100,101]. In ALD, the precursors are injected sequen-
and keeping molar ratios of H2 O/Ti = 25 and H+ /Ti = 0.5. The thin tially in non-overlapping pulses after clearing the reactor of
film over a typical Ti foil is prepared following a sequence involv- previous ones. The coating reaction ends when all the reactive
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 9

sites on the surface are consumed and thus, the maximum mate-

(211)
(101)
rial depositable in a single exposure cycle is determined by the 1000 Anatase
nature of the surface-precursor interaction. This technology allows Rutile
Ti O

(111)
growing the coatings uniformly in small surfaces with high preci- 800 7 13

Intensity (a.u.)

(110)
sion by increasing the number of ALD cycles. Only the synthesis of

(002)
TiO2 thin films by ALD and its application to wastewater treatment 600

(301)
(211)
has been reported by Heikkila et al. [100], who used Ti(OMe)4 and

(1-21)

(1 4 13)
water as precursors injected into the reactor at 10 mbar, 325 ◦ C and

(101)
400

(220)

(112)
with pulses of 0.5 s. The coating obtained were highly stable and

(-125)
(2 0 12)

(200)
(020)
(-130)

(1-41)
(101)
presented good adherence to the substrate. 200
Finally, thermal spray coating technologies allow the synthe-
sis of high quality and reliable TiO2 coatings with great control of 0
physical properties like thickness, porosity, roughness and hard- 10 20 30 40 50 60 70 80
ness [102]. Atmospheric plasma spray (APS) is the most common 22 θ/ (degree
degree )
and versatile thermal spray process, where the coating is produced
using a nano or micropowder precursor that is partially or totally Fig. 6. XRD spectrum with the corresponding crystallographic planes for each phase
melted during the deposition process by the plasma temperature of the TiO2 thin-film photoanode prepared by atmospheric plasma spray and com-
between 6000 and 15,000 ◦ C. Typically, Ar is used to generate the posed of 29% rutile, 9% anatase and 62% Ti7 O13 non-stoichiometric phase.
plasma superheated in a DC arc and the powder feedstock intro- Source: Reproduced with permission from ref. [31].
duced in it is accelerated at high temperature toward the substrate
by the plasma jet. Thus, coating of large surfaces at low operation It has also been reported the formation of coral-like struc-
times is feasible by APS. Garcia-Segura et al. [31] reported for the tures by changing the electrolyte of the cell during Ti anodization.
first time the use of TiO2 coatings onto stainless steel obtained by Liu et al. [111] described the oxidation of Ti plates in an organic
APS (see Fig. 3a) to treat polluted wastewaters by SPEC. The high electrolyte with 1-butyl-3-methylimidazolium tetrafluoroborate
stability and reusability of the coating during consecutive treat- ([BMIM][BF4 ]) at +60 V for 20 min leading to a hierarchical TiO2
ments was well-proven, showing good efficiencies for Acid Orange structure containing nanoholes resembling a coral.
7 decolorization.
3.2.3. Characterization of synthesized photocatalysts
The physical properties of all the synthesized photocatalysts
3.2.2. Nanostructured materials are characterized by means of many techniques. The morphology
Since 2007 TiO2 NTs have been synthesized for PEC purposes. of deposits can be analyzed by SEM, field-emission SEM (FESEM)
The main advantage of these tridimensional and highly ordered and transmission electron microscopy (TEM). From the images
nanostructures compared to thin-films electrodes is due to their obtained by these methods, the distribution and size of TiO2 NTs
higher effective area that allows higher light absorption. can be determined, as described above for the SEM micrograph of
The TiO2 NTs are usually prepared from the anodization of Fig. 3b. For thin-film photocatalysts, the porosity, roughness, adhe-
Ti plates or foils. Before the anodization process, the titanium sion and hardness of deposits can be measured [31]. The chemical
sheets were: (i) ultrasonically degreased with water/acetone, composition of coatings can be assessed by energy dispersive spec-
methanol/acetone or ethanol/acetone mixtures for above 30 min trometry (EDS) and X-ray photoelectron spectroscopy (XPS) and
followed by (ii) mechanical polished with abrasive papers [103] their crystalline structure ascertained by X-ray diffraction (XRD).
or chemical polished with HF/HNO3 /H2 O mixtures [104,105]. The As an example, Fig. 6 presents the XRD pattern of a TiO2 thin film
substrates were subsequently rinsed with water or subjected to coating a stainless steel sheet and prepared by APS [31]. From
another ultrasonic degrease. Once the anodic surface was pre- the crystallographic planes detected, a composition of 29% rutile,
treated, it was anodized in a two-electrode cell with an inert 9% anatase and 62% Ti7 O13 (non-stoichiometric phase) was deter-
cathode of Ni or Pt under vigorous magnetic stirring, employing mined. Furthermore, the crystallites size can be estimated by the
glycerol/water mixtures with H3 PO4 , HF or NH4 F as electrolyte well known Scherrer’s equation:
[89,104–107]. Different anodization conditions have been reported
K
in the literature involving Ecell values from 10 to 60 V for elec- = (13)
␤ cos ␪
trolysis times from 30 min to 17 h in order to obtain TiO2 NTs or
one-dimensionally TiO2 NTAs of various diameters and lengths. The where ␶ refers to the mean size of ordered crystalline domains,
electrochemically synthesized deposits were finally subjected to K is the Scherrer constant, a dimensionless shape factor usually
annealing between 400 and 600 ◦ C for 1–2 h to induce TiO2 crys- with values close to unity, ␭ is the X-ray wavelength, ␤ is the line
tallization and improve the PEC performance [105,108,109]. Fig. 3b broadening at the half value of the maximum intensity in radians
exemplifies the TiO2 NTAs obtained by this procedure by anodizing and ␪ is the Bragg angle in degrees.
at +30 V for 50 h [89]. The optical properties of coatings are usually evaluated by
A similar electrochemical procedure is utilized for the synthesis diffuse reflectance spectroscopy (DRS) using, for instance, an
of TiO2 NBs [70]. In contrast, FTO/TiO2 nanorod arrays are prepared UV/Vis/NIR spectrometer. Fig. 7a exemplifies the DRS spectra
by a hydrothermal method involving the deposition of the photo- obtained for bare TiO2 NTs and TiO2 -NTs doped with different
catalyst onto the cleaned FTO-coated glass from a TiCl4 and HCl proportions of B [74]. All doped materials presented a second
solution heated in a muffle at 180 ◦ C for 2 h [72]. absorption shoulder in the visible region, at ␭ > 400 nm, due to
Several authors [110] proposed an alternative sonoelectro- the introduction of B atoms into the TiO2 lattice. The inset of
chemical method for the preparation of TiO2 NTs. The procedure Fig. 7a shows the Tauc’s graph constructed from the DRS data of
is a modification of the electrochemical procedure described above the undoped and a B-doped material, which was utilized to esti-
based on the use of ultrasonic waves instead of a magnetic stirrer. mate the two Ebg values of 3.2 and 2.2 eV, obtained as the intercept
The irradiation with ultrasonic waves during anodization enhances with the X axis corresponding to the energy h␯ irradiated, accord-
the mobility of ions inside the solution for leading to more efficient ing to ref. [74]. These data allowed the proposal of the previously
production of the nanotubes. discussed PEC mechanism of B-TiO2 NTs shown in Fig. 4a.
10 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

Fig. 7. (a) DRS spectra of: (1) bare TiO2 NTs and doped B-TiO2 NTs with (2) 70, (3)
560, (4) 140 and (5) 280 ppm of NaBF4 . The inset shows the Tauc’s plot and Ebg value
for (1) B280ppm–TiO2 NTs and (2) undoped TiO2 NTs. (b) Photocurrent density vs.
potential curves for (1) dark current, (2) TiO2 NTs before annealing, (3) bare TiO2
NTs and B-doped TiO2 NTs with (4) 70, (5) 140, (6) 560 and (7) 280 ppm of NaBF4 .
Electrolyte 0.1 M Na2 SO4 , scan potential rate 10 mV s−1 and UV/Vis irradiation from
Fig. 8. (a) Nyquist plots for TiO2 NTs and Pd/TiO2 NTs photoanodes in the dark and
a 125 W mercury lamp. Adapted with permission from ref. [74].
under 300 W Xe illumination. (b) Mott−Schottky plot at a fixed frequency of 5 kHz
on TiO2 NTs and Pd/TiO2 NTs under illumination.
The electronic properties of photocatalysts can be assessed by Source: Reproduced with permission from ref. [76].

measuring the spontaneous photocurrent as function of the applied


Eanod in an electrolytic cell. An example is depicted in Fig. 7b for the The much smaller arch of the Pd/TiO2 NTs than that of the TiO2
B-doped materials reported in Fig. 7a [74]. The experiments were NTs in the dark and also under illumination indicates a reduction
conducted using a 0.1 M Na2 SO4 solution of pH 6 and by scanning of the recombination of electron/holes pairs by Pd upgrading the
the potential at a rate of 10 mV s−1 . Scarce photocurrent can be electron mobility. From the corresponding equivalent circuit, the
observed in the dark (curve 1) and before annealing (curve 2) under charge capacitance (C) of the photocatalyst can be determined and
UV/Vis illumination, whereas the photoactivity upgraded strongly subsequently, related to Efb and ND by means of the following equa-
for the bare TiO2 NTs (curve 3). The B-doped samples were more tion:
photoactive than the bare one and showed an increasing photocur-
1 2 kT
rent at higher amount of dopant agent from 70 to 280 ppm (curves = [(E − Efb ) − ] (14)
C2 ND e ε0 ε anod e
4, 5 and 7) due to the enhancing photoactivity of the B, whereas
at 560 ppm (curve 6) the photocurrent decayed since the dopant where e is the elemental charge, ␧0 is the permittivity of the
also trapped the electrons photogenerated in the material. In all vacuum; ␧ is the relative permittivity of the semiconductor, k is
cases, the photocurrent rose with increasing Eanod because a grad- the Boltzmann constant and T is the absolute temperature. The
ual value higher than Efb was applied and this caused a bending corresponding Mott-Schottky plots for the TiO2 NTs and Pd/TiO2
of the CB and VB bands leading to greater SCL (see Fig. 4a) and NTs under illumination are presented in Fig. 8b, where C values
consequently, photocurrent. were acquired beginning at negative Eanod . Thus, an Efb = −0.294 V
An interesting electrochemical method to determine the Efb was found for the former material, a value more negative than
value of the photocatalyst and its electron carrier density (ND ) is −0.137 V determined for the latter one that facilitates the electron
the electrochemical impedance spectroscopy (EIS). The interfacial transfer at any positive Eanod . This was also reflected in the lower
properties between the solution and the photocatalyst are typi- ND = 2.05×1021 cm−3 for TiO2 NTs compared with 8.55 × 1021 cm−3
cally studied from frequencies of 105 to 0.1 Hz using amplitudes of for Pd/TiO2 NTs that is obtained, indicating a faster carrier (electron)
5–10 mV at open circuit. Fig. 8a shows the Nyquist plots (imag- transfer for the latter, which presupposes a best PEC performance.
inary impedance vs real impedance) obtained for TiO2 NTs and The structural characteristics of the photoanodes can affect the
Pd/TiO2 NTs composite in the dark and under 300 W Xe light [76]. photoelectrocatalytic efficiency. Parameters such coating thickness
The semicircles observed are characteristics of a charge transfer or roughness should not be miss considered. However, the study
process and its diameter is equal to the charge transfer resistance. of the influence of these structural parameters has been barely
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 11

attended in the literature to the difficulties to modify and con- a quartz glass tank reactor, an UVC light source and a potentiostat
trol these parameters during the anode synthesis. For instance, [113]. Note that this kind of high energetic light can also photolyze
the effect of the film thickness on the PEC performance could be directly the organic pollutants. The tank reactor contained a Ti/TiO2
ascribed to two main effects: (i) the light penetration into the photoanode and an inert Cu cathode, both placed in parallel, and the
film to photogenerate electron/hole pairs by Reaction (1) and (ii) reference electrode to control the potential of the photoanode was
the coating conductivity. Obviously both effects are related to the a saturated calomel electrode (SCE). The lamp was positioned verti-
photocatalyst material. According to Hitchman and Tian [99], the cally in a double-walled U-tube outside the reactor, surrounded by
maximum penetration depth of incident irradiation into TiO2 coat- circulating water to decrease the heating effect of the lamp. The UVC
ings is defined by 1/␣, where ␣ is the absorption coefficient of TiO2 irradiation supplied by the lamp was perpendicular to the photo-
at the incident light wavelength. It should be noted that the holes catalyst surface and crossed the quartz glass wall without intensity
generated in the depletion layer width should diffuse to the pho- loss. A porous titanium plate was used to provide dissolved oxygen
toanode surface in order to react with water to release • OH by for enhancing the degradation process by O2 − production via Reac-
Reaction (2) or directly mineralize organics. These holes generated tion (3) and pre-eminently, the mass transport of organics toward
in the minority carrier length, which is the distance that the holes the photocatalyst by the stirring of the solution. On the other hand,
move into a field-free region before recombination from Reaction Fig. 9b shows a similar three-electrode configuration, but with an
(7), may diffuse to the depletion layer boundary and transported to inside thermostated Xe light source [114]. The photocatalyst was a
the photoanode surface. On the other hand, the material resistivity Ti/B-TiO2 NTs, the cathode was an inert Ni sheet and the reference
rises when thickness increases, diminishing the efficient removal of electrode was a SCE.
photopromoted electrons by the external electric circuit [35,112]. A recent paper of Martin de Vidales et al. [115] reported a flow
Thereby, the increase of coating thickness enhances the photogen- PEC system with a three-electrode cell to degrade methyl orange
eration up to a limit, where a further increase diminishes the PEC azo dye from wastewater. Fig. 10a depicts an expanded view of the
performance. undivided electrochemical cell utilized, in which the photoanode
Since photocatalytical processes are directly related to surface was a Ti plate coated with TiO2 NTs and the inert cathode was a
processes, the surface roughness of thin films plays a key role Ti mesh coated with RuOx attached over a glass end for the pas-
on the PEC efficiency [112]. Photoanode materials with higher sage of the UV light [115]. The reference electrode was an Ag/AgCl
superficial roughness quadratic average will then result in higher (3 M KCl), connected to the electrolyte exit of the cell, near the
performance. Consequently, the photocatalyst roughness should be photoanode. The arrangement used is illustrated in Fig. 10b. The
determined in order to be able to compare in equality conditions experiments were made with 1 L of solution, which was introduced
the different catalytic materials response. in the reservoir, thermostated at 15 ◦ C with water bath and recir-
culated through the circuit with a centrifugal pump (batch mode).
3.3. PEC systems After passing through the cell, its flow rate was regulated between
20 and 100 L h−1 with a flowmeter before come back to the reser-
Lab-scale photoelectrochemical devices have been utilized up voir.
to present to check the PEC treatment of organic wastewater. The Fig. 11a and b presents two examples of undivided two-
systems are basically composed of an electrolytic cell with the pho- electrode PEC systems for wastewater remediation [31,36]. Daghrir
tocatalyst, a potentiostat or power source to provide the electrical et al. [37] studied the oxidation of chlortetracycline hydrochlo-
energy and a light source to illuminate the photoanode. Never- ride in aqueous solution using the arrangement of Fig. 11a with
theless, a large variety of PEC systems have been described, with a Ti/TiO2 thin film as the photoanode. The tank reactor was made
different cells and kinds and positions (outside or inside the cell) of acrylic material and a quartz window was disposed on one face of
of the light source, which can be UVC (␭ < 300 nm), UVA (␭ from it to illuminate the photocatalyst with UVC light of ␭max = 254 nm.
320 to about 400 nm) or UV/Vis lamps, as well sunlight that can Apart from the Ti/TiO2 , one cathode of the same surface area
be simulated with a Xe lamp. In most works, any criterion is given was placed in parallel. Different cathode materials such as stain-
about the selection of the PEC system applied and as much, the use less steel, vitreous carbon, graphite and amorphous carbon were
of a visible radiation is argued to justify larger absorption in this utilized. The solution was stirred with a magnetic bar. Fig. 11b
region from photoanodes different from classical TiO2 . This variety illustrates a sketch of the set-up used for Garcia-Segura et al.
of conditions makes impossible a realistic comparison of the PEC [31] to degrade Acid Orange 7 azo dye solutions submitted to
performance reported by different authors over a given organic pol- solar irradiation by SPEC. A 5 cm2 TiO2 photoanode and a 3 cm2
lutant, thereby being needed a vast development of optimized PEC carbon-polytetrafluoroethylene (PTFE) air-diffusion cathode were
systems for their possible scale-up to be applicable to industrial used as electrodes. The TiO2 thin-film material was synthesized
level. over stainless steel by APS technology and the cathode produced
The electrolytic cells can be differentiated by their form and continuously H2 O2 from the two-electron reduction of oxygen from
the number of compartments and electrodes. The most ubiquitous injected air by Reaction (8), upon vigorous magnetic stirring. This
cell is a tank reactor, although flow cells are also utilized. One can cathode acted in the degradation process since the generated H2 O2
then operate with: (i) a one-compartment or undivided cell and can directly attack organic pollutants or its oxidation products
(ii) a two-compartment or divided one, with a separator between and/or produce more potent ROS like • OH by Reaction (9).
the anolyte (anodic solution) and catholyte (cathodic solution). The A limited number of papers have described the use of divided
former arrangement is usually preferred for avoiding the potential cells for PEC treatment. Fig. 12a exemplifies the three-electrode
penalty of the separator of the latter one. In both cases, it is feasible arrangement utilized by Selcuk et al. [116] to study the removal
the use of two or three electrodes. The systems with two electrodes of humic acid. The anolyte of the tank reactor of 100 mL capac-
contain the photoanode and cathode (inert for the degradation pro- ity was equipped with a TiO2 photoanode and a SCE, whereas
cess or not), providing a constant Ecell or janod by a power source. In the catholyte with the same solution volume contained a Pt foil
contrast, the three-electrode cells are equipped with an additional cathode. Both compartments were separated by a Nafion 117 mem-
reference electrode that makes possible the supply of a constant brane and the photocatalyst was illuminated with an outside Xe
Eanod to the photoanode by a potentiostat. lamp through a quartz glass window. An interesting study of Ding
Examples of undivided three-electrode PEC systems are et al. [117] compared the performance of divided and undivided
depicted in Fig. 9. The arrangement of Fig. 9a was composed of two-electrode semi-circular quartz glass cylinder cells to degrade
12 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

Fig. 9. Typical set-up of undivided three-electrode cells for wastewater treatment by PEC. (a) TiO2 thin-film photocatalyst for the degradation of a 20 mg L−1 Acid Blue 7 azo
dye solution with external light (Reproduced with permission from Ref. [113]). (b) Doped B-TiO2 NTs photocatalyst for the degradation of a 20 mg L−1 methyl orange azo dye
solution with inner light (Reproduced with permission from Ref. [114]).

200 mL of rhodamine B dye, as shown Fig. 12b. The photoanode 3.4. Operation parameters
was formed by Bi2 WO6 nanoplates deposited on FTO and the cath-
ode was composed of Fe@Fe2 O3 core-shell nanoparticles supported The degradation of organics from wastewaters during PEC
on activated carbon fiber (ACF). The anolyte and catholyte of the process is monitored from the variation of experimental parame-
divided cell were connected with a saturated KCl salt bridge and ters such as the absorbance at the maximum wavelength (␭max )
the photocatalyst was illuminated with a visible light by means of of the solution, usually for dyes, determined by UV/Vis spec-
an outside 300 W tungsten halogen lamp with ␭ > 420 nm. Fresh air trometry, the concentration of the parent molecule measured by
was injected near the cathode to generate H2 O2 by Reaction (8) for high-performance liquid chromatography (HPLC) and total organic
coupling PEC process with electro-Fenton (EF) one. This occurred carbon (TOC) or chemical oxygen demand (COD) of the solution.
thanks to the iron ions released by the cathode during current cir- The two former parameters assess the destruction of the starting
culation, which originated • OH in the bulk from Fenton’s Reaction pollutant/s, the third one informs about its mineralization (conver-
(15) [5,11,118]: sion into CO2 ) and the latter one evaluates its transformation into
small by-products like short-linear carboxylic acids and oxidizable
inorganic species. The change of these parameters can be expressed
H2 O2 + Fe2+ → Fe3+ + • OH + OH− (15) in terms of:

A
Decolorization = (16)
The coupling of PEC with EF mineralized more largely the dye A0
solution in the catholyte of the divided cell than in the undivided
one as a result of a greater accumulation of H2 O2 in the former A0 − A
Percentage of color removal = × 100 (17)
system because it was partly anodically oxidized in the latter one. A0
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 13

Fig. 10. (a) Expanded view of the undivided electrochemical cell with TiO2 NTs photocatalyst and (b) scheme of the set-up of the flow PEC system used for the removal of
1 L of a 0.25 mM methyl orange azo dye solution at pH 3, 15 ◦ C and flow rate of 100 L h−1 .
Reproduced with permission from ref. [115].

c0 − c where A0 . c0 , TOC0 and COD0 denote the absorbance, concentra-


Percentage of organic removal = × 100 (18)
c0 tion, total organic carbon and chemical oxygen demand at initial
time t, and A, c, TOC and COD are the corresponding values at elec-
TOC0 − TOC trolysis time t. It is necessary to remark the typical confusion in
Percentage of TOC removal = × 100 (19)
TOC0 many papers over the indiscriminate use of Eq. (18) instead of Eq.
(17) for dyes solutions. It is well-known that dyes oxidation origi-
COD0 − COD
Percentage of COD removal = × 100 (20) nates colored by-products that can absorb at the same ␭max as the
COD0
14 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

Fig. 11. Schematic diagram of coupled systems of PEC with generated H2 O2 at the cathode using TiO2 thin-film photocatalysts. Treatment of: (a) 1 L of chlortetracycline
hydrochloride solutions (Reproduced with permission from Ref. [36]). (b) 100 mL of a 15 mg L−1 Acid Orange 7 azo dye solution under direct solar irradiation (Reproduced
with permission from Ref. [31]).

parent molecule and hence, the percentage of color removal of the in the SCL with the consequent reduction of PEC activity and grow-
solution determined from absorbance decay from Eq. (17) increases ing of the energy consumption [86]. The light intensity has then to
over time much more slowly than the decay of dye content found influence largely the degradation by PEC since it regulates the quan-
by HPLC from Eq. (18) [119]. Decolorization results reported in the tity of photogenerated electrons that can be extracted at maximum
literature then need to be carefully analyzed for avoiding such bad Eanod . Nevertheless, most authors do not consider this variable and
interpretation. use a high UV or visible intensity to ensure the highest effectiveness
Several operation variables can modify the above experimen- of PEC.
tal parameters and affect the effectiveness of the PEC treatment of Other important variables are the solution pH, its conductivity
organic pollutants in wastewaters. The most important variables related to the electrolyte concentration and the organic concen-
are the external bias potential and light intensity [13–15,18,20–24]. tration that determines the oxidation ability of ROS and dissolved
As stated above, the application of an increasing Eanod respect to O2 [18,21,22]. The effect of pH in PC is explained from the point
Efb accelerates the pass of photogenerated electrons to the exter- of zero charge (PZC) corresponding to the pH where the surface
nal circuit causing a bending of the CB and VB of the photocatalyst charge of the photocatalyst is zero, a value approximately 6 for
with the consequent formation of a SCL. Nevertheless, it is found TiO2 , and the adsorption of organics below and above it [18]. In
that the degradation process of organics is enhanced up to a maxi- contrast, the PZC is not valid in PEC because the surface charge
mum Eanod when the majority of photoexcited electrons are drawn becomes strongly positive upon a given Eanod and the degradation
inside the semiconductor and extracted at the back contact, con- process is limited by the generated oxidizing ROS due to the low
ditions under which the reaction kinetics is only limited by the recombination of electron/hole pairs. The best pH then has to be
photon flux. At superior Eanod values, the width of SCL can exceed experimentally determined in each case since it depends on the
the thickness of the photocatalyst and the charge is redistributed molecule nature and its interaction with the charged photocata-
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 15

Fig. 12. (a) Set-up of a divided PEC system with a TiO2 photocatalyst for treating 100 mL in the photoanode (anolyte) compartment of a 25 mg L−1 humic acid solution
(Reproduced with permission from Ref. [116]). (b) Schematic diagrams of divided (left) and undivided (right) PEC semi-cylindrical cells with a Bi2 WO6 /FTO photoanode and
a Fe@Fe2 O3 /ACF cathode for H2 O2 generation to degrade 10.44 ␮M rhodamine B dye solutions. In both systems, PEC was coupled with electro-Fenton (EF) (Reproduced with
permission from Ref. [117]).

lyst. On the other hand, Na2 SO4 is usually added to the treated and energy costs have been scarcely determined for the PEC treat-
solution for upgrading its conductivity in order to obtain a higher ment of organic pollutants [82,86,107,120,121]. More research over
janod at a given Eanod , improving the PEC effectiveness by a larger these points should be made in order to corroborate the viability of
ROS production. A sufficient amount of dissolved O2 also promotes PEC for future scale-up at industrial level.
oxidizing conditions in PEC, because it helps to avoid the recombi-
nation of photogenerated electron/hole pairs by acting as electron
acceptor from Reaction (3) to form more quantity of ROS as illus- 4. Destruction of organic pollutants by PEC
trated in Figs. 2 and 5. The effect of the temperature and the stirring
rate or flow rate, which determines the reaction rate and the mass The application of PEC to wastewater remediation has been
transfer of reactants/products toward/from the electrodes, respec- preferentially centered on the treatment of dyes, chemicals and
tively, has been reported by a limited number of authors in PEC, pharmaceuticals spiked in synthetic aqueous solutions, usually
despite their basic role in the EAOPs treatments of wastewaters containing Na2 SO4 as electrolyte. Unfortunately, less attention has
[11,118]. been pay on the remediation of real wastewaters due to the com-
It should be noteworthy that the elucidation of the oxidation plexity of their nature and composition. This section is devoted
reaction of organics, the changes in toxicity and biodegradability to show and discuss the main results and advances found for the
treatments of all these kinds of organic pollutants, which will be
16 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

separately analyzed for a better remark of their degradation char- 1.2


acteristics.
1.0
4.1. Dyes

Decolorization
0.8
As highlighted in Fig. 1c, synthetic dyes solutions represent the
major number of wastewaters treated by PEC. Their fast decoloriza- 0.6
tion upon the action of generated ROS can be easily monitored
by UV/Vis spectrometry, even for very low dye concentrations,
0.4
and this allows checking the performance of this technology.
Apart from the large variety of dyes tested, synthesized photoan-
0.2
odes such as TiO2 thin films [31,90,93,95,113,122–136], TiO2 NTs
[69,106,108,115,137–143], doped TiO2 [72,74,114,129,144–146],
0.0
composites with TiO2 [38,76,79,80,84,85,147–158] and other pho- 0 30 60 90 120 150 180 210
tocatalytic materials [52,53,59,60,65,66,117,159–166] have been
Time (min)
utilized in these studies. The main reason for preparing such high
number of photocatalysts is the search of materials with larger abil- Fig. 13. Change of decolorization with electrolysis time for 36 mL of 50 mg L−1 of
ity to photogenerate electron/hole pairs than TiO2 thin films upon methyl orange azo dye in 0.01 M Na2 SO4 using an undivided cell with Ti/TiO2 NTAs
sunlight irradiation. photocatalyst at Eanod = +0.75 V (Ag/AgCl) upon 15 W UVC light. ( ) Direct photolysis
(DP) without photoanode, ( ) photocatalysis (PC) without applied current, ( )
electrochemical oxidation (EO) without UVC illumination and ( ) PEC. Adapted
4.1.1. TiO2 photoanodes with permission from Ref. [106].
Table 1 collects the percentages of color and TOC removals
obtained for different dye solutions treated by PEC under selected
conditions using undivided cells with either TiO2 thin-film or NTs rapidly the dye. The same synergistic behavior for PEC has been
photocatalyst (see Figs. 9–11). A look of these data allows con- shown for Acid Orange 7 [31], Remazol Brilliant Orange 3R [90],
cluding that artificial UVC (␭max = 254 nm), UV (from 180 to ca. rhodamine B [93] and orange G [136].
600 nm) and sometimes Xe lamps with high intensity were pref- The operation variables most studied in dye degradation are
erentially used to irradiate the photocatalyst. Exposure to UVA the solution pH, the applied Eanod , the dye concentration, the elec-
(␭max = 360 nm) lamps [90] and direct sunlight [31,95] has been trolyte concentration and the kind of electrolyte used. Fig. 14
utilized, as well. Dye concentrations were typically low (normally illustrates the effect of the above operation variables over the
< 25 mg L−1 ) due to the small oxidation ability of PEC, because low degradation of 350 mL of methyl orange solutions in sulfate
Eanod values, associated with small photocurrents, were utilized medium by PEC with a TiO2 thin film upon 350 W UV illumination
to preserve the fragile photoactivity of the TiO2 photoanode. For [129]. Fig. 14a evidences that the best pH for the process was 2.0,
this reason, operation in batch mode with long electrolysis times with percentages of color removal slightly higher than that of pH 3.5
of 2–3 h were commonly applied to achieve more than 70% of color and much more superior to those of pH 5.2. This behavior is diffi-
removal (see Table 1). cult to explain and differs from the optimum pH values reported
Comparative tests for given dyes using both, TiO2 thin-film by other authors using Ti/TiO2 thin films. Fu et al. [113] also
and TiO2 NTs photocatalyst under the same experimental PEC described an optimum acidic pH = 3.4 with 90% color removal for
conditions have been scarcely reported. The superior mechani- 20 mg L−1 of Acid Blue 7 in sulfate medium using the cell of Fig. 9a at
cal integrative structure and photoactivity of the nanoestructured Eanod = +0.68 V (SCE) upon 11 W UVC for 1 h (see Table 1), whereas
material is assumed by most authors without verification. The Garcia-Segura et al. [31] found a pH = 7.0 as optimal for 100% decol-
Zanoni’s group [138,140] described that after 240 min of PEF treat- orization of 100 mL of 15 mg L−1 of Acid Orange 7 solutions in
ment at Eanod = +1.0 V (Ag/AgCl) under UV irradiation, 0.05 mM 0.05 M Na2 SO4 using the cell of Fig. 11b at janod = 1.0 mA cm−2 and
Disperse Red 1, Disperse Orange 1 and Disperse Red 13 in chloride 35 ◦ C after 120 min of electrolysis under direct sunlight. Conversely,
solutions undergone 87–97% mineralization using a TiO2 NTs pho- Li et al. [93] reported an increasing decolorization from pH 2.0–10.0
toanode prepared by anodization, values much higher than 54–60% when 50 mL of a 0.015 mM rhodamine B solution in 0.10 M Na2 SO4
obtained for a TiO2 thin-film one prepared by sol-gel. These authors was treated at Eanod = +0.50 V (SCE) exposed to 30 W UVA, attain-
also found a strong reduction of the toxicity of the treated dye ing the best color removal of 95% after 70 min of electrolysis at
solutions, thereby showing that PEC is a very interesting method the higher pH = 10.0. This discrepant behavior for pH can be mainly
for wastewater remediation. This is due to the conversion of the related to: (i) the different systems, photocatalysts and operation
toxic aromatic dye and its aromatic intermediates into biodegrad- conditions utilized that change the amounts of • OH produced from
able products like short-linear carboxylic acids. The formation of photogenerated holes and the interaction of organics with the pho-
tartaric, succinic, acetic and oxamic acids and their removal over tocatalyst and (ii) the different colored products formed from each
electrolysis time has been reported for the SPEC process of Acid dye that interfere the absorbance measurements for decoloriza-
Orange 7 [31]. tion [31]. When 0.50 M NaCl was used as electrolyte, Zanoni et al.
Several papers have described the best performance of PEC over [90] reported a maximum TOC removal of 28% for 30 mL of 0.05 mM
individual processes for dye decolorization. Fig. 13 exemplifies the Remazol Brilliant Orange 3R at pH 6.0 as the anolyte of a divided cell
decolorization abatement for 50 mg L−1 methyl orange in 0.01 M operating at Eanod = +1.0 V (SCE) upon 450 W Xe–Hg arc illumina-
Na2 SO4 by different methods [106]. Individual processes like direct tion for 30 min. The superior degradation at this pH explored within
photolysis (DP) under 15 W UVC, EO at Eanod = +0.75 V (Ag/AgCl) and the range 2–11 could be explained by the synergistic action of the
PC under 15 W UVC led to a poor decolorization lower than 20%. strong oxidant • OH formed from Reaction (2) and active chlorine
In contrast, the PEC treatment yielded 97% color removal, demon- species like HClO and Cl• produced from Cl− oxidation by Reactions
strating that it was much more potent and synergistic to EO and (21)–(23), respectively [129].
PC because the larger separation of the photoexcited electron/hole 2Cl− → Cl2 + 2 e− (21)
pairs under the applied Eanod generated much greater amount of
− +
• OH from photogenerated holes via Reaction (2) that oxidized more Cl2 + H2 O → HClO + Cl + H (22)
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 17

Table 1
Per cent of color and TOC removals for selected PEC treatments of dyes using undivided cells with TiO2 photoanodes.

Dye co (mg L−1 ) Experimental conditions % color removal % TOC Ref.


decay

Acid Blue 7 20 Solution of pH 3.4, Eanod = +0.68 V (SCE), 90 −c [113]


11 W UVC for 1 ha
Methyl orange 10 165 mL of solution, Eanodl = +0.6 V, 100 −c [122]
20 125 W UV for 2 ha 78
Triazo leather dye 10 70 mL of solution with 0.1 M KCl, 35 ◦ C, 98 −c [127]
Eanod = +0.8 V (Ag/AgCl), 125 W UV for (pH 6.14)
8 ha 72
(pH 3.35)
Methyl orange 20 350 mL of solution with 0.1 M Na2 SO4 , 97 −c [129]
pH 2.0, Eanod = +0.75 V (Ag/AgCl), 350 W
UV for 150 mina
Reactive Orange 16 100 70 mL of solution with 0.1 M KCl, 100 −c [134]
Reactive Black 5 100 pH 3.35, 30 ◦ C, Ecell = 5 V, 125 W UV 92 −c
Reactive Red RB 133 100 for 5 ha 70 −c
Rhodamine B 5 20 mL of solution with 0.1 M 53 −c [135]
phosphate, pH 7, Eanod = +0.8 V
(Ag/AgCl), 500 W Xe for 2 ha
Rhodamine B 0.01d 45 mL of solution with 0.1 M Na2 SO4 , 72 −c [69]
Eanod = +0.6 V (SCE), 60 W visible light
for 3 hb
Methyl orange 50 36 mL of solution with 0.01 M Na2 SO4 , 98 −c [106]
Eanod = +0.75 V, 15 W UVC for 3 hb
Methyl orange 0.25d 1 L of solution with 0.1 M Na2 SO4 , pH 3, 51 −c [115]
15 ◦ C, Eanod = +1.50 V (Ag/AgCl), flow
rate 100 L h−1 , 20.5 A of UV for
150 minb
Drimaren Red 243 X −6BN 25 2 L of solution with 4.2 mM KCl, 93 −c [142]
Eanod = +1.50 V (Ag/AgCl), 30 W UVC for
1 hb
a
TiO2 thin film.
b
TiO2 NTs.
c
Not determined.
d
Initial concentration in mM.

HClO + hv → Cl• + • OH (23) tions (21)–(23), and (iii) H2 O2 from SO4 2− oxidation via Reactions
(25)–(26), which could generate extra • OH by Reaction (9).
NO3 − + H2 O + h␯ → NO2 • + • OH + OH− (24)
Fig. 14b highlights that increasing Eanod from +0.2 to +1.0 V
2− 2− −
(Ag/AgCl) accelerated the degradation process of methyl orange 2SO4 + h␯ → S2 O8 + 2e (25)
because more current was gradually generated enhancing the sep- 2− −
S2 O8 + 2H2 O → 2HSO4 + H2 O2 (26)
aration of electrons and holes. However, the degradation rate
decreased for Eanod = +2.0 V (Ag/AgCl) suggesting a redistribution The action of these additional oxidants differed for other dyes.
of charges in SCL with greater electron/hole recombination despite For instance, Zanoni et al. [90] described a large rise in TOC removal
the circulation of a higher current through the electrodes. Low in the order: Na2 SO4 < NaNO3 < KCl when 0.05 mM Remazol Bril-
optimum Eanod values have also been shown for the degradation liant Orange 3R solutions in 0.50 M of such electrolytes at pH 6.0
of several dyes using TiO2 thin-films or TiO2 NTs photoanodes as the anolyte of a divided cell were treated at Eanod = +1.0 V (SCE).
in different systems, varying between +0.6 and +1.2 V (SCE) In this case, the dye molecule was more effectively attacked by
[90,93,106,113,122,127,129], although up to +1.50 V (Ag/AgCl) has generated active chlorine species than by • OH.
been reported [115]. The positive effect of temperature from 20 to 35 ◦ C and of
The expected decay of the percentage of color removal with light intensity from 44 to 132 mW m−2 on the decolorization of
increasing methyl orange concentration from 10 to 30 mg L−1 is 25 mg L−1 Acid Black 1 dye using an undivided cell are shown in
depicted in Fig. 14c. This is a typical behavior of EAOPs and can be Fig. 15a and b, respectively [108]. Macedo et al. [127] described a
accounted for by the destruction of lower percentage of organics similar behavior respect to temperature for the color abatement of
with increasing their concentration since a similar amount of • OH a 10 mg L−1 triazo leather dye solution in 0.10 M KCl of pH 3.35 and
oxidants is produced from photogenerated holes via Reaction (2) 6.14 using a TiO2 thin film at Eanod of +0.8 and +1.2 V (SCE) under
[5]. 125 W UV operating between 21 and 35 ◦ C. This trend is expected
Fig. 14d shows that increasing Na2 SO4 concentration up to by the concomitant enhancement of the mass transport of reac-
0.10 M as bakcground electrolyte caused greater degradation rate tants toward/from the photoanode and the increase in rate of all
of the dye as a result of the rise of the solution conductivity. giving chemical and electrochemical reactions involved in the decoloriza-
rise to a higher current that produced more holes and consequently, tion process. For much higher intensities between 5 and 20 W m−2
more quantity of • OH to oxidize methyl orange. Finally, Fig. 14e of UVC illumination, Fu et al. [113] found that 16 W m−2 was opti-
evidences that the decolorization rate for different 0.10 M elec- mal for the decolorization of a 20 mg L−1 Acid Blue 7 solution using
trolytes slightly increased in the order: NaNO3 < KCl < Na2 SO4 . This the cell of Fig. 9a at Eanod = +0.68 V (SCE), indicating a saturation of
sequence could be tentatively ascribed to the growing conductiv- photoexcited electrons in the VB of a TiO2 photocatalyst at high
ity of the solutions at pH 2.0, with little influence of the additional enough light intensity.
production of: (i) • OH from NO3 − photolysis by Reaction (24), (ii) Martin de Vidales et al. [115] used the flow cell of Fig. 10 to
lesser oxidant active chlorine species from Cl− oxidation by Reac- degrade 1 L of 0.25 mM methyl orange in 1 M Na2 SO4 at pH 3, 15 ◦ C
18 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

a b
100 100

80 80

60 60

40 40 0.2 V
pH 2.0 0.6 V
pH 3.5 1.0 V
20 pH 5.2 20 1.5 V

0 0
0 30 60 90 120 150 180 0 30 60 90 120 150 180
c d
100 100
% Color removal

80 80

60 60

40 -1 40
10 mg L
L-1 0M
-1
20 mg L
L-1 0.05 M
-1
20 30 mg L
L-1 20 0.10 M

0 0
0 30 60 90 120 150 180 0 30 60 90 120 150 180
e
100

80

60

40
NaNO
NaNO33
KCl
KCl
20 Na SO
Na2SO4
2 4

0
0 30 60 90 120 150 180
Time (min)

Fig. 14. Percentage of color removal vs. electrolysis time for the degradation of 350 mL of methyl orange solutions in sulfate medium by PEC using an undivided cell with
Ti/TiO2 thin-film photocatalyst upon 350 W UV illumination. (a) Effect of pH, 20 mg L−1 methyl orange, 0.10 M Na2 SO4 , Eanod = +1.0 V (Ag/AgCl), (b) influence of the applied Eanod ,
20 mg L−1 methyl orange, 0.10 M Na2 SO4 , pH 2.0, (c) effect of methyl orange content, 0.10 M Na2 SO4 , pH 2.0, Eanod = +1.0 V (Ag/AgCl). In (d) influence of sulfate concentration
and (e) effect of different 0.10 M electrolytes for 20 mg L−1 methyl orange, pH 2.0 and Eanod = +1.0 V (Ag/AgCl). Adapted with permission from ref. [129].

and Eanod = +1.50 V (Ag/AgCl) by varying the flow rate. After 150 min ysis of Fe(OH)2+ species (the most stable Fe(III) species at pH 3.0)
of electrolysis, the color removal grew from 20% at 20 L h−1 to 51% at from Reaction (27) and (ii) the photolysis of Fe(III) complexes with
100 L h−1 . At the higher flow rate the process achieved the best effi- final carboxylic acids by the general Reaction (28) [5,11]:
ciency because it was controlled by the mass-transport of reactants
Fe(OH)2+ + h → Fe2+ + • OH (27)
and the oxidation products were rapidly removed to the solution
leaving the surface photoactive sites available for further organic Fe(OOCR) 2+
+ h → Fe 2+
+ CO2 + R• (28)
oxidation. This condition can also be easily attained in tank reactors
with vigorous stirring (mechanic, magnetic or by bubbling O2 ) of The coupled PEC/PEF process led to total decolorization in
the solution, as shown in the systems of Figs. 9–12. 300 min with 74% mineralization. This method was much more
The coupling of PEC with cathodically generated H2 O2 has powerful than the PEC treatment alone, without Fe2+ addition,
also been explored aiming to enhance its degradation power which yielded only 50% color removal at the same time.
[31,123,126,136]. Xie and Li [126] proposed the use of an undivided Another proposed coupled system involves the combination of
tank reactor similar to that of Fig. 9a equipped with a Ti/TiO2 mesh PEC and EF in a divided tank reactor similar to that of Fig. 12b [123].
photocatalyst upon outside 8 W UVA and a reticulated vitreous car- The anolyte contained 100 mL of 30 mg L−1 of Acid Scarlet 3R dye
bon cathode fed with 120 mL min−1 of air. The anode mesh allowed in 0.02 M Na2 SO4 and a Ti/TiO2 thin-film electrode illuminated by a
the illumination of 30 mL of a 30 mg L−1 Orange G dye solution in 125 W UV light. The catholyte was separated of the anolyte by a salt
0.01 M Na2 SO4 with 0.15 mM Fe2+ of pH 3.0 and room temperature. bridge and contained 50 mL of the same solution adjusted to pH 3.0
Under the application of a cathodic potential (Ecat ) of −0.71 V (SCE), with 0.036 mM Fe2+ and a graphite electrode fed with an air flow of
H2 O2 was continuously produced at the cathode by Reaction (8) and 2.5 L min−1 . Operating at Ecat = −0.66 V (SCE) for 70 min, 60% color
oxidizing • OH radicals were pre-eminently formed from photogen- removal was achieved in the anolyte through PEC process, whereas
erated holes by Reaction (2) and in the bulk from Fenton’s Reaction a much higher decolorization of 92% was found in the catholyte by
(15). Furthermore, the UVA illumination of the solution allowed the EF. This work evidences that a coupled PEC/EF process with a “dual
application of the photoelectro-Fenton (PEF) process involving: (i) cell” can upgrade the dye degradation compared with single PEC.
the additional generation of • OH with Fe2+ regeneration by photol- Recently, Bessegato et al. [143] showed the combination of PEC
and ozonation for improving the destruction of Acid Yellow 1 dye
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 19

1.2 a
a 100
1.0 20 ºC
25 ºC 80

% Color removal
30 ºC
Decolorization

0.8
35 ºC 60
0.6
40

0.4
20

0.2 0
0 30 60 90 120 150
0.0 Time (min)
0 10 20 30 40 50 60 70
Time (min) b
100
1.2
b 80

% TOC removal
-2
44 mW
44 mW m--2
m
1.0 -2
88 mW m
88 60
-2
132 mW m
132
Decolorization

0.8
40
0.6
20

0.4 0
0 30 60 90 120 150
0.2 Time (min)

Fig. 16. Percentage of (a) color removal at ␭ = 393 nm and (b) TOC removal for the
0.0
0 10 20 30 40 50 60 70 degradation of 500 mL of a 100 mg L−1 Acid Yellow 1 with 0.1 M Na2 SO4 solution at
pH 2.0 and 25 ◦ C in an undivided cell. Method: ( ) EO with Ti/B-TiO2 NTs, ( ) DP,
Time (min) ( ) PC with Ti/TiO2 NTs, ( ) PC with Ti/B-TiO2 NTs, ( ) PEC with Ti/TiO2 NTs and
( ) PEC with Ti/B-TiO2 NTs. Irradiation with 125 W UV/Vis light and Eanod = +1.2 V
Fig. 15. Decolorization vs. electrolysis time for the PEC treatment of 100 mL of a (Ag/AgCl) in EO and PEC. Adapted with permission from ref. [74].
25 mg L−1 Acid Black 1 dye solution using an undivided cell with Ti/TiO2 NTs photo-
catalyst irradiated by UVC light (Eanod was not specified). (a) Effect of temperature
at light intensity of 44 mW m−2 and (b) influence of light intensity at 25 ◦ C. Adapted the much smaller recombination of the electron/hole pairs under
with permission from Ref. [105]. Eanod = +1.2 V (Ag/AgCl) in PEC yielded quicker destruction of organ-
ics. Under these conditions, as well as in PC, the doped B-TiO2 NTs
photocatalyst gave better degradation than TiO2 NTs one. This is
solutions up to 100 mg L−1 in 0.01 M Na2 SO4 at pH 3.0. A TiO2 NTs
due to the reduction of the Ebg value of the former material from
photoanode in an annular bubble reactor operating at Ecell = 2.0 V
3.2 to 2.2 eV due to the B-doping level (see Fig. 4a). The better results
with 36 W UVB irradiation in the presence of 1.25 × 10−4 mol min−1
of 100% color removal and 93% TOC decay were obtained with an
of O3 was employed. Total decolorization and mineralization were
amount of 280 ppm of B as doping agent.
attained in 20 and 60 min for the coupled PEC/O3 , respectively,
A similar enhancement of PEC with dopant B has been reported
which presented much higher oxidation ability than individual pro-
by Su et al. [114], who used the arrangement of Fig. 9b to remark the
cesses.
beneficial of B-TiO2 NTs photocatalysts under Xe illumination over
methyl orange decolorization. While improved PEC processes for
4.1.2. Doped TiO2 and composites with TiO2 dyes were also found for dopant Cu2+ [72] and codopants W and
A high number of modified TiO2 photocatalysts has been synthe- N [145], any effect was observed using either Fe3+ [144] or V, Ce
sized with the aim of upgrade the performance of the PEC process and F [129] as doping agents of TiO2 thin films. On the other hand,
of dyes using visible light as radiation source. Most of published positive PEC performance of different dyes has been described by
articles have only detailed the comparative decolorization and using composites of TiO2 with metals [76,147] and other materials
mineralization of these pollutants in sulfate medium between the [79,80,84,85,148–157].
modified TiO2 and the pristine material. Fig. 16a and b exemplify The percentage of color and TOC removals for several dyes
the change of the percentage of color and TOC removals with elec- degraded by PEC using different doped TiO2 and composites with
trolysis time, respectively, for different treatments of 100 mg L−1 of TiO2 photoanodes under selected conditions are summarized in
solutions of a nitronaphthol dye like Acid Yellow 1 in 0.1 M Na2 SO4 Table 2. The smaller Ebg values of these photocatalysts compared to
at pH 2 and 25 ◦ C [74]. The cell was a tank reactor with an arrange- that of TiO2 allowed their best irradiation with visible light alterna-
ment similar to that of Fig. 9b equipped with either a TiO2 NTs tively to UVC and UVA. The data of Table 2 show the use of potent
or a doped B-TiO2 photoanode under 125 W UV/Vis irradiation by Xe lamps in many cases for photo-excitation, which simulate in
applying an Eanod = +1.2 V (Ag/AgCl) in EO and PEC [74]. EO in the the laboratory the role of sunlight. Higher Eanod values than those
dark practically did not yield any removal of the dye, indicating used for TiO2 in Table 1 can be observed in Table 2 because of
that it was not oxidized at the anode. A large decolorization with the larger stability of the modified materials. Good color removals,
poor TOC reduction can be observed under DP, suggesting that the usually >80%, were obtained, although electrolysis times as long
dye and some of its by-products were directly and slowly pho- as 2–3 h were needed operating in batch mode. Similarly, TOC
tolyzed by the incident light. The decay of both parameters was removals >70% were found under these conditions when dye con-
strongly enhanced by PC pre-eminently by the larger oxidation tents <50 mg L−1 were degraded. Greater contents required much
effectiveness of • OH radicals formed from Reaction (2). As expected, longer electrolysis times. For instance, a 105 mg L−1 rhodamine 6G
20 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

Table 2
Percentage of color and TOC removals for selected degradations of dyes by PEC using undivided cells with doped TiO2 and composites with TiO2 photoanodes.

Dye co (mg L−1 ) Experimental conditions % color removal % TOC decay Ref.

Acid Yellow 1 50 500 mL of solution with 0.01 M Na2 SO4 , pH 2, 100 93 [74]
25 ◦ C, B-TiO2 NTs photocatalyst, Eanod = +1.2 V
(Ag/AgCl), 125 W UV for 2 h
Methyl orange 20 500 mL of solution with 0.1 M Na2 SO4 , 25 ◦ C, 92 −a [114]
B-TiO2 NTs photocatalyst, Eanod = +2.0 V (SCE),
300 W Xe for 2 h
Methyl orange 10 50 mL of solution with 0.2 M Na2 S, CdS/TiO2 89 −a [84]
NTAs photocatalyst, Eanod = +0.5 V (SCE), 300 W
Xe for 3 h
Methyl orange 10 200 mL of solution with 0.1 M NaCl, pH 2.7, 97 −a [79]
SiO2 /TiO2 photocatalyst, Eanod = +0.8 V (SCE),
9 W UVC for 30 min
Methylene blue 5 3 mL of solution with 0.5 M Na2 SO4 , pH 6, 25 ◦ C, 63 −a [72]
Cu-TiO2 NTAs photocatalyst, Eanod = +1.0 V
(Ag/AgCl), 105 W visible light for 4 h
Methylene blue 4 10 mL of solution with 0.1 M KCl, 89 −a [154]
graphene/TiO2 /carbon cloth, Eanod = +0.9 V
(SCE), 150 W Xe for 160 min
Methylene blue 4 Each solution with 0.05 M Na2 SO4 , Pd/TiO2 NTs 100 −a [76]
photocatalyst, Eanod = +0.5 V (Ag/AgCl), 300 W
Rhodamine B 2.5 Xe for 3 h 84 −a
Rhodamine B 10 Solution with 0.01 M Na2 SO4 , −a 82 [148]
TiO2 /carbon/Al2 O3 membrane, Eanod = +1.2 V
(SCE), 500 W Xe for 2 h
Rhodamine 6G 105 250 mL of solution with 0.1 M Na2 SO4 , 25 ◦ C, 100 69 [152]
W-Ti NTAs photocatalyst, Eanod = +0.3 V (SCE),
80 W UV for 8 h
a
Not determined.

solution needed up to 8 h to achieve 69% TOC reduction [152] (see The degradation of the dye (1) is initiated with its desulfonation
Table 2). All these results evidence again the low oxidation ability and hydroxylation to give the compound 2. This intermediate can
of PEC, useful for low organics loads. then undergo: (i) successive hydroxylation to compounds 3 and 4
Few studies have described the influence of pH [74,79] and or (ii) C N cleavage to form the naphthalene products 5 and 6 with
applied Eanod [74,79,114] over the degradation of dyes by PEC. desulfonation. Compound 5 evolves to the benzoic acid derivative
The analysis of their decolorization process revealed that they 7, whereas the oxidation of 6 yields the benzene compounds 8 and
always obeyed a pseudo-first-order kinetics owing to the attack 9, along with the naphthalene compound 10, which can also be pro-
of a constant • OH content produced from photogenerated holes via duced from 4. Degradation of 7 leads to the sulfobenzene product 11
Reaction (2) [72,76,84,114,156]. Fraga et al. [80] treated 30 mL of a that is oxidized to the p-benzoquinone derivative 12, also formed
0.033 mM Basic Red 51 solution in 0.1 M Na2 SO4 of pH 2.0 using an from compounds 8-10. The ring opening of benzenic intermediates
undivided cell with a WO3 /TiO2 bicomposite photoanode and a Pt gives the aliphatic carboxylic acids 13 and 14, which are then trans-
cathode. After 120 min of electrolysis at janod = 1.25 mA cm−2 upon formed into the ultimate oxalic (15) and formic (16) acids. These
irradiation of the photocatalyst with UV (280–400 nm) and visible acids form Fe(III)-oxalate and Fe(III)-formate complexes that are
(420–630 nm) light, total decolorization was always attained with photolyzed to CO2 with Fe2+ regeneration via Reaction (28).
94% and 83% TOC abatement, respectively. Using comparable PC, The coupling of membrane filtration with PEC has been designed
only 25% and 30% color removal and 11% and 7% TOC decay were by Wang et al. [148]. The membrane was constructed deposit-
found. These results corroborate again the superiority of PEC over ing graphitic carbon layer and TiO2 nanoparticles on an Al2 O3
PC and the feasible sustainable solar-assisted remediation of water membrane support. When this TiO2 /carbon/Al2 O3 membrane was
bodies contaminated with dyes. irradiated with a 500 W Xe lamp and an Eanod = +1.2 V (SCE) was
The effectiveness of the coupled PEC/PEF process was assessed provided for 2 h, 82% TOC reduction was obtained for a 10 mg L−1
by Almeida et al. [38] for 500 mL of 85 mg L−1 orange G solutions rhodamine B solution with 0.01 M Na2 SO4 flowing through the
in 0.05 M Na2 SO4 at optimum pH 3.0 and 25 ◦ C using an undi- membrane (see Table 2). This removal was 1.3 or 3 times higher
vided cell like of Fig. 9a by applying 16.67 mA cm−2 . This method than that found for filtration with light irradiation (using PC pro-
was conducted with a Pt/TiO2 NTs photoanode irradiated with cess) or filtration alone. The coupled PEC/membrane became more
a 80 W UV/Vis lamp and a carbon-PTFE air-diffusion cathode for efficient thanks to the greater mineralization of the dye with the
continuous H2 O2 electrogeneration. It was compared with single larger amounts of • OH formed from Reaction (2) upon the action of
PEC using an alternative Pt cathode and with EF using an alter- the applied Eanod .
native Pt anode. A 0.5 mM Fe2+ concentration was added to the
solution for PEC/PEF and EF treatments. The dye disappeared com- 4.1.3. Other photocatalytic materials
pletely in 16, 12 and 6 min for PEC, EF and PEC/PEF, respectively, Semiconductor materials like WO3 and composites
always obeying a pseudo-first-order kinetics. After 190 min, TOC [52,53,160,161,166], Bi2 WO6 [117], ZnO [162], BDD-ZnWO4
was reduced by 80%, 87% and 97% for the above degradations. [159], ␤-PbO2 [65], BiVO4 [66], BiPO4 [164,165] and ␣-Fe2 O3
The coupled PEC/PEF process then showed higher oxidation ability [59,60,163], with an appropriate Ebg to photo-excite electron/hole
than PEC alone. LC–MS/MS and HPLC analysis of treated solutions pairs similarly to Reaction (1) for TiO2 , have been checked for
revealed the formation of 12 aromatic intermediates and 4 short- the remediation of dye solutions by PEC. Table 3 collects the
linear carboxylic acids. From these products, the reaction sequence degradation parameters obtained for several dye treatments with
for orange G mineralization of Fig. 17 was proposed [35], where the the above photocatalysts using undivided cells under selected
main oxidizing agents are • OH produced from Reaction (2) and (15). conditions. As can be seen, UVC, UV, visible light and Xe lamps
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 21

Fig. 17. Proposed reaction sequence for orange G mineralization by the coupled PEC/PEF process using an undivided cell with Pt/TiO2 NTs photoanode upon 80 W UV/Vis
irradiation and a carbon-PTFE air-diffusion cathode. Adapted with permission from ref. [38].

were successfully applied to decolorize low dye concentrations in general, quite similar to those applied with undoped TiO2 (see
(<25 mg L−1 ) for long times, usually between 2 and 3 h, due to the Table 1) and doped TiO2 and composites with TiO2 (see Table 2).
low generation of • OH from photogenerated holes. Note that the Most of the above papers corroborated the superior oxida-
experimental conditions used for the photocatalysts of Table 3 are, tion power of PEC compared to individual processes like DP, EO
and PC [60,65,159,161,163,164]. For example, Chatchai et al. [161]
22 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

Table 3
Per cent of color and TOC removals for selected PEC treatments of dyes in undivided cells with other photocatalytic materials.

Dye co (mg L−1 ) Experimental conditions % color % TOC Ref.


removal decay

Direct Red 80 5 200 mL of solution with 5 g L−1 Na2 SO4 , 25 ◦ C, 88 12 [65]


25 ␤-PbO2 photocatalyst, I = 5 mA, 7 W UVC for 6 h 72 17
X-3B 20 60 mL of solution with 0.1 M Na2 SO4 , 97 −a [159]
ZnWO4 /BDD photocatalyst, Eanod = +2.0 V (SCE),
20 W UV for 3 h
Basic Red 51 0.01b 30 mL of solution with 0.1 M Na2 SO4 , pH 2, 100 63 [160]
W/WO3 photocatalyst, janod = 1.25 mA cm−2 ,
150 W Xe for 2 h
Rhodamine B 0.1 b 200 mL of solution, WO3 photocatalyst, 98 81 [166]
Eanod = +1.5 V (SCE), direct sunlight for 160 min
Methylene blue 5 Solution with 0.1 M Na2 SO4 , pH 6.5, 83 −a [161]
WO3 /BiVO4 photocatalyst, Eanod = +0.2 V
(Ag/AgCl), 500 W Xe (> 420 nm) for 2 h
Methylene blue 0.01b 5 mL of solution with 0.1 M Na2 SO4 , BiVO4 51 −a [66]
photocatalyst, Eanod = +1.4 V (Ag/AgCl), 50 W
visible light for 40 min
Methylene blue 5 25 mL of solution with 0.01 M Na2 SO4 , pH 7, 78 −a [163]
␣-Fe2 O3 photocatalyst, Eanod = +0.6 V (SCE),
350 W Xe for 105 min
Methyl orange 5 50 mL of solution with 0.1 M Na2 SO4 , ␣-Fe2 O3 79 −a [59]
photocatalyst, Eanod = +1.0 V (SCE), 300 W
visible light for 2 h
Methyl blue 10 Solution with 0.1 M Na2 SO4 , pH 7, BiPO4 80 −a [164]
photocatalyst, Eanod = +3.0 V (SCE), 11 W UVC
for 5 h
a
Not determined.
b
Initial concentration in mM.

reported the progressive decay of the more intense visible band of 6


methyl orange at ␭max = 662 nm up to 83% after 2 h of PEC treatment a
Initial specific rate (10 M cm s )
-2 -1

of a 5 mg L−1 dye solution in 0.1 M Na2 SO4 at pH 6.5 using an undi- 5


vided cell equipped with a WO3 /BiVO4 composite photocatalyst
at Eanod = +0.2 V (Ag/AgCl) under a 500 W Xe (>420 nm) irradiation. 4
-9

This process was much more powerful than PC alone, which only
yielded about 26% color removal, whereas the decolorization by DP 3
and EO at Eanod = +0.2 V (Ag/AgCl) was rather insignificant. These
results confirm an excellent separation of the electron/hole pairs 2
upon a bias potential for a WO3 /BiVO4 composite.
Nevertheless, little information has been given over the influ- 1
ence of operation variables like pH [52,53,161], applied Eanod
[52,53,65,159,164], light [52,160,161], dye concentration [52,65] 0
and background electrolytes [52,66]. In their pioneering works 0.0 0.2 0.4 0.6 0.8 1.0 1.2
with a WO3 , Luo and Hepel [52,53] studied the degradation of [NaCl] (M)
the naphthol blue black diazo dye in different electrolytes with a
conventional undivided three-electrode cell like of Fig. 9a under 4
b
Initial specific rate (10 M cm s )

irradiation with an outside 500 W UV/Vis light. For a dye concen-


-2 -1

tration of 8.8 × 10−5 M with 0.5 M electrolytes and Eanod = +1.08 V


(SCE), the decolorization of the PEC process was enhanced in the 3
sequence: Na2 SO4 < NaNO3 < NaClO4 < NaCl. The best performance
-9

in 0.5 M NaCl can be attributed to the very fast attack of active chlo-
rine species formed from Reactions (21)–(23) compared to that of 2
• OH produced from photogenerated holes by Reaction (2), which

are the only oxidants expected in the stable 0.5 M NaClO4 . The faster
decolorization in 0.5 M NaNO3 than in the latter one can then be 1
related to the generation of more amounts of • OH via the photolytic
reaction (24), whereas the lower oxidation ability of S2 O8 2− pro-
duced from Reaction (25) could justify the slowest loss in color 0
achieved in 0.5 M Na2 SO4 . The oxidation action of active chlorine 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
species was confirmed by the linear dependence between the initial [dye] (10 M)
-4

specific rate for decolorization and NaCl concentration depicted in


Fig. 18a [52], assessed under optimum conditions of 8.8 × 10−5 M Fig. 18. (a) Initial specific rate vs. NaCl concentration for the decolorization of
dye concentration and Eanod = +1.08 V (SCE). It was also found that 8.8 × 10−5 M naphthol blue black diazo dye solutions by PEC using an undivided tank
reactor with WO3 photocatalyst upon 500 W UV/Vis light at Eanod = +1.08 V (SCE). (b)
the initial specific rate in 0.5 M NaCl was quite similar in the pH
Initial specific rate vs. naphthol blue black content for the PEC decolorization of dye
range 1.0-4.5, whereas it dropped progressively up to pH 12.0 by solutions in 0.50 M NaCl by illuminating the WO3 photocatalyst with 500 W of ( )
the loss of oxidation power of active chlorine species due to the UV/Vis and ( ) visible light at Eanod = +1.08 V (SCE). Adapted with permission from
formation of the weaker oxidant ClO− . Moreover, the illumination ref. [52].
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 23

40 1.2 1.2
a a
1.0 1.0
30
EO
% TOC removal

0.8 0.8

Decolorization

TOC / TOC
PEC

20 0.6 0.6

0.4 0.4

0
10
0.2 0.2

0 0.0 0.0
-1
5 mg L 5 mA
-1
25 mg L 5 mA 25 mg L
-1
50 mA 0 20 40 60 80 100 120 140
Time (min)
1.2 1.2 1.2
b
b
1.0 1.0
1.0
Decolorization

0.8 0.8

Decolorization

TOC / TOC
0.8
0.6 0.6
0.6
0.4 0.4

0
0.4 0.2 0.2

0.2 0.0 0.0


0 1 2 3 4 5 6 0 20 40 60 80 100 120 140
Time (h) Time (min)

Fig. 19. (a) Percentage of TOC removal for EO and PEC treatments of 200 mL of 5 and Fig. 20. Time-course of decolorization decay (in red) and normalized TOC removal
25 mg L−1 Direct Blue 80 dye solutions in 5 g L−1 Na2 SO4 at 25 ◦ C using an undivided (in blue) in the (a) undivided and (b) divided semi-cylindrical tank reactors of Fig. 12b
cell equipped with ␤-PbO2 photocatalyst upon 7 W UVC illumination for 6 h at 5 mA for the treatments of 200 mL of a 1.044 × 10−5 M rhodamine B solution in 0.05 M
and 4 h at 50 mA (adapted with permission from Ref. [65]). (b) Decolorization vs. Na2 SO4 of pH 6.2 upon illumination of the Bi2 WO6 /FTO photoanode with 300 W
electrolysis time of a 10 mg L−1 methyl blue dye solution with 0.1 M Na2 SO4 at pH 7 visible light and at janod = 0.1 mA cm−2 . (a) Systems: ( ) Bi2 WO6 /FTO|Pt, ( )
by ( ) EO at Eanod = +3.0 V (SCE), ( ) PC and PEC at Eanod of ( ) +1.0 V, ( ) +2.0 V, Pt|Fe@Fe2 O3 /ACF and ( ) Bi2 WO6 /FTO|Fe@Fe2 O3 /ACF. (b) Catholyte degrada-
( ) +3.0 V and ( ) +4.0 V (SCE) in an undivided cell with a BiPO4 photocatalyst tion in ( ) Pt||Fe@Fe2 O3 /ACF and ( ) Bi2 WO6 /FTO||Fe@Fe2 O3 /ACF. Adapted
under 11 W UVC irradiation (adapted with permission from Ref. [164]). with permission from ref. [117].

with UV/Vis light gave much better decolorization rate than visible with ability to release iron ions and electrogenerate H2 O2 to pro-
light, as can be seen in Fig. 18b, indicating that the photo-excitation duce • OH from Fenton’s Reaction (15), as stated above. A very
of electrons in WO3 was mainly due to UV light, despite its low Ebg low janod of 0.1 mA cm−2 was used, thus avoiding the formation of
• OH from water oxidation by Reaction (10). In the undivided cell,
of 2.5–2.7 eV (see subsection 3.1.2).
The combined effect of dye concentration and applied current Fig. 20a illustrates that the decolorization and TOC decay for the
on TOC decay for the EO and PEC processes of 200 mL of Direct Blue coupled PEC/EF process (Bi2 WO6 /FTO|Fe@Fe2 O3 /ACF system) were
80 dye solutions in 5 g L−1 Na2 SO4 at 25 ◦ C using an undivided cell greater than the sum of those of PEC (Bi2 WO6 /FTO|Pt system) and
with a ␤-PbO2 photocatalyst upon 7 W UVC light was analyzed by EF (Pt|Fe@Fe2 O3 /ACF system) [117]. The synergistic process of the
Florencio et al. [65]. Fig. 19a evidences a strong improvement of coupled treatment can then be accounted for by the larger gener-
PEC over EO at the greater dye content of 25 mg L−1 and higher ation of • OH from both Reactions (2) and (15). The production of
• OH was enhanced when the dye degradation was performed in
current of 25 mA, which can be ascribed to a large enhancement
of the separation electron/hole pairs upon the action of external the catholyte of the divided cell, with better performance using a
potential. On the other hand, the high optimum Eanod value that can Bi2 WO6 /FTO photocatalyst than a Pt anode, as shown in Fig. 20b.
be applied to a BiPO4 photocatalyst was evidenced by Zeng et al. These results are rather doubtful, because if a janod = 0.1 mA cm−2
[164] when exploring the decolorization of a 10 mg L−1 methyl blue was applied in both cases, the same degradation parameters should
solution with 0.1 M Na2 SO4 at pH 7 using 11 W UVC illumination. be expected for the EF process taking place in the catholyte. In fact,
Fig. 19b shows that the rise of Eanod from 0 to +3 V (SCE) enhanced the degradation of the catholyte in the divided cell can only be
the removal of the dye up to a maximum 80% color removal at 5 h associated with EF but not with a coupled PEC/EF since the photo-
(see Table 3). Further increase to +4 V (SCE) caused the redistribu- generated • OH cannot attack the organic load of the solution.
tion of charges in SCL with greater electron/hole recombination and
only 67% color removal was obtained at the same time. For all these 4.2. Chemicals
trials, a pseudo-first-order kinetics was always found.
Ding et al. [117] examined the effectiveness of the coupled A high number of wastewaters containing chemicals have been
PEC/EF treatment of rhodamine B in 0.05 M Na2 SO4 at neutral pH treated by PEC aiming to show the interest of this technology for
6.2 in the undivided and divided semi-cylindrical tank reactors their remediation (see Fig. 1c). Most of these compounds are widely
of Fig. 12b using a Bi2 WO6 /FTO photocatalyst exposed to 300 W used toxic and biorefractory industrial products such as anilines,
visible illumination. The cathode was a Fe@Fe2 O3 /ACF electrode phenols, bisphenol A and carboxylic acids. Photocatalysts including
24 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

TiO2 thin films and NTs [30,44,89,99,109,116,167–173], doped TiO2 a


[174–176], composites with TiO2 [70,75,82,86,177–187] and other 100
materials [60,64,67,188–190] under different Eanod values and illu-
mination sources have been utilized in these studies, as will be 80

% TOC removal
described below.
60
4.2.1. TiO2 and composites with TiO2
Table 4 collects the degradation characteristics of several chem-
icals destroyed by PEC under the action of TiO2 and composites 40
with TiO2 photoanodes under selected conditions. High chemi-
cal removals (>92%) along with excellent TOC reductions can be 20
observed in most cases using any kind of photoanode illuminated
with UV, visible and/or simulated solar light. Long times from 1 0
to 5 h were again required in view of the relatively low amount of 0 30 60 90 120 150 180 210
oxidizing agents originated during the PEC treatment of low con- Time (min)
tents of chemicals. In some cases, the decay of solution COD was
determined. This parameter reflects the amounts of organic and 100
inorganic species that can be oxidized by dichromate ion, whereas
b
TOC accounts for by the amount of organic carbon present in the

[Active chlorine] (mg L )


80

-1
solution. The COD/TOC ratio varies with the compound checked
and, for example, attains a value of 2.38 for phenol and only 0.18
for oxalic acid. Thus, the COD value for wastewaters with aromatic 60
compounds decreases more rapidly than TOC because of the for-
mation of final short-linear aliphatic carboxylic acids like formic 40
and oxalic [5]. A COD removal up to 95% has been reported for
the destruction of a 10 mM formic acid solution by PEC using Ag-
TiO2 [177] photocatalysts under 500 W visible light at Eanod = +0.8 V 20
(SCE) for 60 min. On the other hand, it was found that 50 mL with
a 817 mg L−1 solution of this acid underwent a lower COD reduc- 0
tion of 63.8% after 60 min of PEC treatment in a packed-bed reactor 2.0 4.0 6.0 8.0 10.0
containing a granular Ti/TiO2 thin film exposed to 500 W visible pH
illumination at Ecell = 30.0 V [168]. These findings suggest the fea-
sibility of a large mineralization of organic pollutants because the Fig. 21. (a) Percentage of TOC removal with time for 250 mL of a 5.0×10−6 M 4,4 -
oxydianiline solution in 0.1 M Na2 SO4 of pH 2.0 in an undivided cell under 150 W
PEC process is potent enough to remove final carboxylic acids to
UVA light. ( ) DP, ( ) PC with Ti/TiO2 thin film, ( ) PC with Ti/TiO2 NTAs, ( )
large extent. PEC with Ti/TiO2 thin film at Eanod = +1.5 V (Ag/AgCl) and ( ) PEC with Ti/TiO2
Similarly to dyes, most works demonstrate the best per- NTAs at Eanod = +1.5 V (Ag/AgCl) (Reproduced with permission from Ref. [89]). (b)
formance of PEC to destroy chemicals from wastewaters than pH-dependence of active chlorine concentration produced in 250 mL of a 0.1 M
individual processes. The superiority of TiO2 NTs [89], doped TiO2 NaCl in an undivided cell with a Ti/TiO2 thin-film photocatalyst upon 125 W UV
(315–400 nm) irradiation for 30 min (Reproduced with permission from Ref. [170]).
[174,175] and composites with TiO2 [177,178,185] over thin films
of this material has also been well-proven. Fig. 21a shows the com-
parative TOC removal for a 5.0 × 10−6 M 4,4 -oxydianiline solution and 2,4-dichlorophenol [185] from addition of radical scavengers
with 0.1 M Na2 SO4 at pH 2.0 by several methods and photocatalysts like tert-butanol (• OH scavenger), oxalate (hole scavenger) and p-

[89]. DP and PC with Ti/TiO2 thin film and NTAs under 150 W UVA benzoquinone (O2 − scavenger).
light only led to 40–50% mineralization in 180 min. At that time, The effect of pH on the PEC performance was also dependent on
TOC was more largely reduced by 78% using PEC with a Ti/TiO2 the molecule nature. In most cases, the best removal of chemicals
thin-film photocatalyst at Eanod = +1.5 V (Ag/AgCl), whereas total was obtained in an acidic pH range 1–4 [89,116,169,172,175,183],
mineralization was attained in a shorter time of 120 min when the although for aniline [44] and dodecyl-benzenesulfonate [174] the
photocatalyst was a Ti/TiO2 NTAs material. This work shows clearly optimum pH values were close to 10 and 6.2, respectively. For 0.05-
the enhancement of the separation of the photogenerated elec- 0.10 M NaCl, Fraga et al. [170] confirmed the larger production of
tron/holes by highly ordered NTAs with much greater photoactive active chlorine species at pH 4.0 using an undivided reactor with a
area than conventional thin films. Ti/TiO2 thin-film photocatalyst upon 125 W UV light, as can be seen
Contradictory relative oxidation abilities have been described in Fig. 21b. They obtained 95% TOC reduction of a 50 ␮g L−1 micro-
referring to the electrolytes used in synthetic solutions. While an cystin solution at this pH using janod = 30 mA cm−2 for 180 min.
increasing PEC performance in the order NaCl < NaNO3 < Na2 SO4 Chai et al. [82] presented an interesting study on the degrada-
has been found for 4,4 -oxydianiline [89], humic acid [116] and p- tion of a 200 mg L−1 p-nitrophenol solution in 0.1 M Na2 SO4 in an
nitrophenol [167], the sequence NaNO3 < Na2 SO4 < NaCl has been undivided cell similar to that of Fig. 9a to check the performance of
obtained for microcystin-LR [183]. Moreover, Jorge et al. [167] a SnO2 /TiO2 NTs composite under 300 W UV illumination by sev-
reported a superior performance of HClO4 instead of the above eral treatments. A high content of this pollutant was utilized to
three electrolytes, whereas Daskalaki et al. [172] obtained a bet- better determine the changes in biodegradability and relative tox-
ter removal of bisphenol A using NaCl instead HClO4 . This different icity index of the treated solution, and the evolution of detected
behavior can be associated with the different reactivity of parent intermediates. The biodegradability was determined as the ratio
molecules and its oxidation products with generated ROS and/or of biochemical oxygen demand at 5 days (BOD5 ) and COD, need-
active chlorine species, as pointed out above. Several studies in ing a threshold value of 0.4 for biological post-treatment. The
sulfate medium have confirmed the oxidation preponderance of relative toxicity index with respect to the starting sample was cal-
• OH formed from reaction (2) over O • − produced from reac- culated from the change in luminescence of the marine bacterium
2
tion (3) during the PEC treatment of tetrabromobisphenol A [70] Vibrio fisheri. Fig. 22a reveals a very small p-nitrophenol concentra-
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 25

Table 4
Percentage of chemical and TOC removals for selected PEC processes of several organics in undivided cells with TiO2 and composites with TiO2 photocatalyts.

Chemical co (mg L−1 ) Experimental conditions % chemical removal % TOC decay Ref.

4,4 -oxydianiline 0.1a 250 mL of solution with 0.1 M Na2 SO4 , pH 6.0, 92 31 [89]
TiO2 NTs photocatalyst, Eanod = +1.5 V
(Ag/AgCl), 150 W UVA for 180 min
Phenol 9.4 40 mL of solution with 0.1 M Na2 SO4 , pH 6, 95 18 [186]
bifunctional TiO2 NTs and Ni-Sb-SnO2
photocatalyst, Eanod = +2.0 V (SCE), 150 W Xe for
100 min
p-Nitrophenol 200 100 mL of solution with 0.1 M Na2 SO4 , 25 ◦ C, 98 91 [81]
SnO2 /TiO2 NTs photocatalyst, Eanod = +1.5 V
(SCE), 300 W UV for 4 h
p-Nitrophenol 5 60 mL of solution, pH 2, PANI/Cr-TiO2 NTs 99 76c [175]
photocatalyst, Eanod = +0.8 V (SCE), 15 W UVC
for 2 h
Bisphenol A 0.2a 60 mL of solution with 0.1 M NaCl, pH 5, 98 −b [172]
ITO/TiO2 photocatalyst, janod = 0.32 mA cm−2 ,
150 W Xe for 75 min
Bisphenol A 15 160 mL of solution, pH 4.5, 97 −b [86]
TiO2 /graphene/Cu2 O, Eanod = +0.5 V (SCE),
500 W Xe for 150 min
Tetrabromobisphenol A 5 35 mL of solution with 0.5 M Na2 SO4 , Au/TiO2 97 37 [70]
NBs photocatalyst, janod = 20 mA cm−2 , 35 W
visible light for 100 min
Humic acid 25 100 mL of solution with 0.0125 M NaCl, pH 3, 100 83 [169]
Ti/TiO2 photocatalyst, Eanod = +1.0 V (SCE),
450 W Xe for 150 min
Microcystin-LR 1 40 mL of solution with 0.01 M NaNO3 , 92 78 [183]
TiO2 /AgCl/Ag photocatalyst, Eanod = +0.6 V
(SCE), 60 W visible light for 5 h
a
Initial concentration in mM.
b
Not determined.
c
COD decay.

225
a 100 b
200
[p-Nitrophenol] (mg L )
-1

175
80
% TOC removal

150
125 60
100
75 40

50
20
25
0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
1.2 5
c d
1.0
4
Relative toxicity index

0.8
BOD /COD

3
0.6
5

2
0.4
1
0.2

0.0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
Time (h)

Fig. 22. Change of (a) p-nitrophenol concentration, (b) percentage of TOC removal, (c) biodegradability and (d) relative toxicity index with time for the treatment of 100 mL
of 200 mg L−1 pollutant solution with 0.1 M Na2 SO4 in an undivided cell for 4 h. Method: ( ) PC with Ti/SnO2 , ( ) PC with SnO2 /TiO2 NTs, ( ) EO with TiO2 NTs, ( ) EO
with SnO2 /TiO2 NTs and ( ) PEC with SnO2 /TiO2 NTs. Eanod = +1.5 V (SCE) and illumination with 300 W UV light.
Source: Reproduced with permission from ref. [81].

tion abatement by PC processes with both Ti/SnO2 and SnO2 /TiO2 NTs at the same bias anodic potential in 4 h. The p-nitrophenol con-
NTs photocatalysts [81], which was strongly enhanced by EO at centration decay always obeyed a pseudo-first-order decay. The
Eanod = +1.5 V (SCE), more pronounced for the composite material, same trend can be observed in Fig. 22b for TOC removal, where the
due to the larger generation of • OH via Reaction (10). The addi- most powerful PEC process attained a larger mineralization of 91%.
tional photogeneration of ROS by Reactions (2)–(5) allowed the Under EO and PEC conditions, the biodegradability of the solution
total disappearance of the parent molecule by PEC with SnO2 /TiO2 upgraded to values >0.4, whereas the relative toxicity index was
26 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

reduced to practically zero using these methods with the SnO2 /TiO2 a
NTs composite. However, low contents of by-products like hydro- 1.0
quinone, p-benzoquinone and maleic, fumaric and oxalic acids still
0.8
remained in the final solution treated by PEC. These results corrob-
orate the potential applicability of this procedure at industrial level,
0.6

0
A/A
being feasible its coupling with a cheaper biological post-treatment
to produce reusable water.
0.4
For the PEC process in an undivided cell, optimum Eanod values of
+0.8 V (SCE) for p-nitrophenol with PANI/Cr-TiO2 NTs [175], +1.0 V 0.2
(SCE) for p-nitrophenol with TiO2 thin film [167] and dodecyl-
benzenesulfonate with doped W-TiO2 [174], +1.5 V (Ag/AgCl) for 0.0
4,4 -oxydianiline with Ti/TiO2 NTAs [89] and +2.0 V (SCE) for phe- 0 30 60 90 120 150 180 210 240 270
nol with bifunctional TiO2 NTs and Ni-Sb-SnO2 photocatalyst [186], Time (min)
along with optimum janod = 0.32 mA cm−2 for bisphenol A with b
ITO/TiO2 [172], have been found. Moreover, the expected decrease 1.0
of the pseudo-first-order rate constant for organic decay and per-
0.8
centage of TOC removal with increasing chemical concentration
due to the presence of more organic matter has been confirmed by
0.6

0
A/A
several authors [29,44,89,167,174,175]. Similar results have been
described for divided cells [116,169,178]. Selcuk et al. [116,169]
0.4
utilized the divided cell of Fig. 12a to degrade 100 mL of 25 mg L−1
humic acid in sulfate or chloride media, contained in the anolyte, 0.2
using a Ti/TiO2 photocatalyst exposed to 450 W Xe irradiation. They
found for 0.0125 M NaCl, optimum operation conditions of pH 3 and 0.0
Eanod = +1.0 V (SCE) for total pollutant removal with 83% TOC decay 0 30 60 90 120 150 180
after 150 min of electrolysis (see Table 4), a value quite similar to Time (min)
that obtained in 0.01 M Na2 SO4 at the same conditions.
Fig. 23. Variation of the relative absorbance decay at ␭ max = 276 nm with elec-
The best performance of the PEC process with increasing light trolysis time for 160 mL of 15 mg L−1 bisphenol A at pH 4.5 under (a) 85 and (b)
intensity has also been confirmed by Yang et al. [86] from the 100 mW cm−2 Xe illumination. Method: ( ) DP, ( ) PC with TiO2 /Cu2 O mesh,
decay of the absorbance of the UV band at ␭max = 276 nm of bisphe- ( ) PC with TiO2 /graphene/Cu2 O mesh, ( ) PEC with TiO2 /graphite/Cu2 O mesh at
nol A. Fig. 23a and b evidence the large stability of 15 mg L−1 of Eanod = +0.5 V (SCE) and ( ) PEC with TiO2 /graphite/Cu2 O mesh and 50 mM H2 O2
addition at Eanod = +0.5 V (SCE). Adapted with permission from ref. [86].
this chemical at pH 4.5 upon DP when it was illuminated with
a Xe lamp at 85 and 100 mW cm−2 , respectively. The use of PC
with a TiO2 /Cu2 O mesh caused a notable abatement in bisphenol little information was given over the optimization of operation vari-
A concentration, which was significantly improved by applying a ables. On the other hand, Nissen et al. [189] checked the removal of
TiO2 /graphene/Cu2 O mesh due to the effective electron/hole sep- 60 mL of 40 mg L−1 of 2,4-dichlorophenol in water (without elec-
aration, as shown in Fig. 5. The bisphenol A removal was strongly trolyte) of pH 5.5 in the anolyte of a divided H-cell by supplying
upgraded by PEC at Eanod = +0.5 V (SCE), as expected by a greater 35 W visible irradiation to the immersed WO3 photocatalyst. Since
extraction of electrons from the CB of Cu2 O enhancing the for- a very low photocurrent of 7.5 ␮A was applied owing to the low
mation of holes at the VB of TiO2 with higher formation of • OH conductivity of the medium, only 53% of the content of this pollu-
radicals from Reaction (2). The removal of this chemical was largely tant was removed after 48 h of treatment.
improved by the production of more quantities of these radicals Phenolic compounds can be directly oxidized at the Bi2 WO6
upon addition of 50 mM H2 O2 via Reaction (9). Fig. 23b reveals a photocatalyst surface, e.g., at +0.8 V (SCE) for 4-chlorophenol [188],
97% reduction of bisphenol A after 150 min of PEC (see Table 4), and hence, a higher Eanod upgrades their direct destruction. The first
which decreased to 90 min upon 50 mM H2 O2 addition. Compar- step in EO of phenol or chlorinated phenols involves the formation
ison of Fig. 23a and b clearly shows a more rapid destruction of of a phenoxy radical, which can undergo either radical–radical or
bisphenol at higher light intensity for both PC and PEC processes radical–substrate coupling to give polymeric products that remain
because of the greater photogneration of electron/hole pairs. adsorbed on the electrode surface and inhibit the process, espe-
It has also been reported the positive action of H2 O2 on PEC cially in alkaline medium where their deprotonated forms are
when it was electrogenerated on site at an ACF/PTFE [106] or involved. In contrast, the photogenerated ROS species in PEC can
boron-doped diamond (BDD) cathode [172] via Reaction (8). With activate the electrode surface and promote pollutant oxidation.
the latter cathode, for example, the bisphenol A concentration in Fig. 24a shows that the quicker removal of 4-chlorophenol con-
60 mL of a 0.2 mM solution in 0.1 M NaCl at pH 5 was reduced centration at pH 6.5 with Bi2 WO6 exposed to 500 W Xe lamp was
by 98% using an ITO/TiO2 photocatalyst under 150 W Xe light at achieved at Eanod = +2.0 V (SCE) [188], where the separation of the
janod = 0.32 mA cm−2 for 75 min (see Table 4), whereas the alterna- electron/hole pairs was maximal. At higher Eanod values up to +4.0 V
tive use of a Zr cathode (no H2 O2 production) only led to 42% of (SCE) the SCL became gradually larger and the photogenerated
concentration removal. electrons were redistributed, facilitating its recombination with
holes and making the process more inefficient. Fig. 24b illustrates
4.2.2. Other photoanodes the gradual drop in 2,4-dichlorophenol content abatement for the
The effective application of ␣-Fe2 O3 [60], Sb-SnO2 [64], Bi2 WO6 above PEC treatment at optimum Eanod = +2.0 V (SCE) when pH var-
[67,188], WO3 [189] and ZnO [190] semiconductors to the destruc- ied from an alkaline medium of pH 8.5 to an acidic one of 4.3
tion of phenolic compounds by PEC has also been reported. The [188]. This behavior can be related to a progressive larger adsorp-
good degradation results obtained for such treatments in undi- tion of the molecule that favors its reaction with oxidizing agents
vided cells after long irradiation time with UVA, UVC or Xe lamps formed at the photoanode surface, although the smaller forma-
are summarized in Table 5. All these studies described again the tion of inhibitory polymeric films with decreasing pH could also
enhancement of PEC performance over individual processes, but contribute to the better efficiency of the process in acidic medium.
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 27

Table 5
Per cent of chemical and TOC removals for selected PEC treatments of phenols using undivided cells with other photocatalytic materials.

Chemical co (mg L−1 ) Experimental conditions % chemical removal % TOC decay Ref.

Phenol 20 100 mL (without electrolyte), SnO2 -Sb2 O4 100 85 [64]


photocatalyst, Ecell = 2.0 V, 250 W UVA for 2 h
Phenol 10 15 mL of solution with 0.5 M Na2 SO4 , Bi2 WO6 78 65 [67]
photocatalyst, Eanod = +0.6 V (SCE), 500 W Xe for
3h
p-Nitrophenol 0.1a 50 mL of solution with 0.1 M Na2 SO4 , pH 3.0, 91 74b [190]
ZnO photocatalyst, Eanod = +1.4 V (SCE), 15 W
UVC for 3 h
p-Nitrophenol 0.05a 50 mL of solution with 0.1 M Na2 SO4 , ␣-Fe2 O3 47 −c [59]
photocatalyst, Eanod = +1.0 V (SCE), 300 W
visible light for 4 h
4-Chlorophenol 10 100 mL of solution with 0.5 M Na2 SO4 , pH 4.3, 68 65 [188]
Bi2 WO6 photocatalyst, Eanod = +2.0 V (SCE),
500 W Xe for 12 h
a
Initial concentration in mM.
b
COD decay.
c
Not determined.

60 4.3. Pharmaceuticals
a
50 Despite the social alarm over the presence of pharmaceu-
% 4-chlorophenol removal

ticals in the aquatic environment, their feasible degradation


40 by PEC has only been assessed for few compounds in the
last years [37,81,88,107,191–200]. The trials were pre-eminently
30 made with TiO2 photoanodes, although composites with TiO2
[81,191,198,199] and new Ti compounds [88] have been utilized as
20 well. Table 6 summarizes the percentages of drug and TOC decays
achieved by this technique for some compounds under selected
10 conditions using undivided cells with TiO2 thin-film and NTs photo-
catalyst. As seen previously, UVA, UVC and Xe lamps were applied to
0 treat the solutions. Low drug concentrations (usually ≤ 20 mg L−1 )
0 1 2 3 4 5
were used to check the oxidation ability of the technique and good
E (V vs. SCE)
anod per cent of drug removals (regularly >80%) were obtained for times
12 longer than 1 h.
b Most authors confirmed the expected larger oxidation ability
10 of PEC to remove drugs over individual processes. The effect of
[4-Chlorophenol] (mg L )
-1

some operation variables such as pH [37,192,196–198], applied


8 Eanod or I [192,196–198] and drug [37,196–198] and electrolyte
[197] concentrations over the process performance was assessed.
6 The kinetics for drug decay [37,88,191,192,196,197], the detection
of oxidation products [37,107,191,198], the drop in toxicity during
4 treatment [37,107,196], economical assessment [37,191] and even
the use of scavengers [197,199] has been reported as well.
2 An interesting work of Daghrir et al. [37] showed that the
optimum PEC treatment of 1 L of 0.025–0.230 mg L−1 chlortetra-
0 cycline solutions in 0.050 M Na2 SO4 took place at pH close to 6,
0 3 6 9 12 15
at I = 0.39 A under 6.9 mW cm−2 UVC irradiation using the arrange-
Time (h) ment of Fig. 11a with a vitreous carbon cathode as compared to
stainless steel, graphite and amorphous carbon ones. Under these
Fig. 24. (a) Percentage of 4-chlorophenol removal vs. anodic bias potential for the
degradation by PEC of 100 mL of 10 mg L−1 organic solution with 0.5 M Na2 SO4 at pH conditions, 98% drug decay and 67% TOC removal after 120 min of
6.5 using an undivided cell like of Fig. 9a with a Bi2 WO6 photoanode under 500 W Xe treatment of a 0.025 mg L−1 solution were obtained (see Table 6),
light for 8 h. (b) 4-Chlorophenol content decay with electrolysis time for the above with similar values in the concentration range tested and an esti-
solution at pH: ( ) 8.5, ( ) 6.5 and ( ) 4.3 and Eanod = +2.0 V (SCE). mated cost of 4.52 US$ m−3 for chemicals plus energy consumption
Source: Reproduced with permission from ref. [188]. [37,193]. The superiority of the vitreous carbon cathode was related
to its ability to produce H2 O2 from Reaction (8), which can be
photolyzed to OH by UVC from Reaction (29), thus enhancing the
Fao et al. [190] found a 91% pollutant reduction with 74% organics removal from • OH generated by Reaction (2).
COD removal after 3 h of PEC treatment of 50 mL of a 0.1 mM p-
nitrophenol solution in 0.1 M Na2 SO4 with a ZnO photocatalyst H2 O2 + hv → 2 • OH (29)
upon 15 W UVC illumination at optimum Eanod = +1.4 V (SCE) and The authors reported a strong reduction of the toxicity of the
pH 3.0 (see Table 5). When the p-nitrophenol concentration grew 0.025 mg L−1 solution after optimum treatment during 120 min
up to 0.5 mM, a gradual loss in its abatement up to 33% was obtained from microtox and microalga tests. Moreover, the primary oxi-
as a result of the larger organic matter treated with similar quan- dation products were detected by LC–MS proposing the reaction
tities of oxidants photogenerated. A pseudo-first-order kinetics for sequence of Fig. 25 [37]. The degradation of chlortetracycline
all concentration decays was always determined. (1) under the main action of OH is initiated by the formation
28 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

Table 6
Per cent of drug and TOC removals for selected PEC treatments of pharmaceuticals using undivided cells with TiO2 photoanodes.

Drug co (mg L−1 ) Experimental conditions % drug removal % TOC decay Ref.

Chlortetracycline 0.025 1 L of solution with 0.05 M Na2 SO4 , 25 ◦ C, pH 98 67 [37]


near 6, I = 0.39 A, 6.9 mW cm−2 UVC for
120 mina
Chlortetracycline 25 1 L of solution with 0.05 M Na2 SO4 , pH 4.8, 74 −c [193]
25 ◦ C, I = 0.39 A, 6.9 mW cm−2 UVC for 120 mina
Tetracycline 10 500 mL of mixed solution with 0.1 M Na2 SO4 , 81 −c [192]
Oxytetracycline 10 pH 7.4, Eanod = +1.0 V (SCE), 15 W UVA for 83 −c
Chlortetracycline 10 180 minb 85 −c
Triclosan 20 50 mL of solution with 0.02 M Na2 SO4 , pH 5.8, 79 60 (1 h) [107]
Eanod = 0 V (SCE), 125 W UVC for 30 minb
Acyclovir 20 0.1 mL of solution with 0.2 M NaNO3 in a 97 −c [194]
thin-layer reactor, pH 5.8, Eanod = +1.0 V
(Ag/AgCl), 10 mWcm−2 UVA for 370 sb
Ofloxacin 40 70 mL of solution with 0.1 M Na2 SO4 , pH 3.0, 100 70 [196]
Eanod = +0.8 V (SCE), 15 W UVC for 35 minb
Sulfamethoxazole 1.3 50 mL of solution with 10 mM NaCl, pH 2.7, 100 − c
[197]
Eanod = +0.5 V (Ag/AgCl), 4 W UVA for 150 mina
Diclofenac 5 80 mL of solution with 0.1 M Na2 SO4 , 100 −c [200]
Eanod = +0.4 V (SCE), 35 W Xe for 10 hb
a
TiO2 thin film.
b
TiO2 NTs.
c
Not determined.

of six derivatives: isochlortetracycline (2) by dechlorination, epi- was degraded in an undivided cell with a Ti/TiO2 thin-film photo-
demeclocycline (3) and 4-epi-N-demethylchlortetracycline (4) by catalyst by applying Eanod = +0.5 V (Ag/AgCl) exposed to 4 W UVA
demethylation, 4-epi-N-dedismethylchlortetracycline (5) by the light [197]. After 80 min of PEC treatment, the sulfamethoxazole
loss of two methyl groups, isochlortetracycline (6) with forma- content dropped to 72%, which was reduced to 60% or 12% by adding
tion of a 5-ring heterocycle and its hydroxylated product 7. Further 2.4 mg L−1 humic acid or 10% methanol, respectively. The two lat-
hydroxylation, demethylation or dedismethylation of 7 yields the ter compounds are well-known scavengers of generated ROS and
compounds 8, 9 and 10, respectively. active chlorine species and then, it react more rapidly with these
Optimum degradation for this antibiotic and its mixture with oxidants diminishing the destruction rate of the antibiotic. The
other tetracyclines in the pH range 6–9 has also been reported decrease of the corresponding pseudo-first-order rate constant is
by Liu et al. [192], whereas acidic media of pH 1–3 were found depicted in Fig. 26b. This figure also reveals a photocurrent density
optimal for the oxidation of ofloxacin [196,198] and sulfamethox- near 0.14 mA cm−2 for the systems with sulfamethoxazole alone
azole [197]. The best performance of chlortetracycline at neutral or humic acid added, but it underwent an unexpected spectular
pH was explained from the higher electrostatic attraction between growing to 0.46 mA cm−2 by adding 10% methanol. The excep-
its zwitterionic form and the positively charged surface of Ti/TiO2 tional enhancement of the photocurrent density in the presence of
[37]. methanol was justified by current doubling that refers to the release
The decay of chlortetracycline [37,192] and sulfamethoxazole of additional electrons from reactive solute-hole intermediates. It
[197] was analyzed from a kinetic equation involving a Langmuir- was hypothesized the oxidation of a methanol molecule by a hole
Hinshelwood model that presupposes a rapid adsorption of the to produce a methoxy radical (CH3 O• ) by Reaction (30), followed
molecule at the photocatalyst surface and the limiting step is the by the oxidation of this radical to formaldehyde with release of
subsequent transformation of the adsorbed species. This model one electron by Reaction (31). Nevertheless, more information is
was found valid for relatively high drug concentrations, but it was needed to confirm this assumption.
reduced to a pseudo-first-order equation at low contents. The lat-
ter equation has been directly applied to explain the abatement of CH3 OH + h+ VB → CH3 O• + H+ (30)
ciprofloxacin [88], paracetamol [191] and ofloxacin [196]. + −
CH3 O• → CH2 O + H + e CB (31)
The reduction of toxicity of the treated solution by PEC is due to
the destruction of the parent molecule and its oxidation products.
This was clearly shown by Liu et al. [107] for the oxidation of 50 mL 4.4. Real wastewaters
of 20 mg L−1 of the bactericide triclosan in 0.02 M Na2 SO4 at natu-
ral pH 5.0 using an undivided cell with a Ti/TiO2 NTs photocatalyst The remediation of a very limited number of real wastewaters by
at Eanod = 0 V (SCE) under 125 W UVC light for 60 min. They deter- PEC has been investigated at lab-scale using several photocatalysts
mined the solution toxicity from the luminescent Photobacterium upon exposure to UV and visible lights. Good color removal and/or
phosphoreum T3 spp. bacterium, expressed as equivalent HgCl2 con- COD abatements at relatively short times have been found for river
tent. The toxicity was largely reduced during the first 10 min of PEC water contaminated with humic acid using Ti/TiO2 thin film [201],
due to the loss of the toxic triclosan. Further, it was increased up sugarcane factory wastewater using ZnO and doped N-ZnO [202],
to 30 min where it reached a quasi-steady value owing to the for- pharmaceutical wastewater using Ni/TiO2 thin fim [203], landfill
mation of the more toxic intermediate 2,7-dichlorodibenzodioxin, leachate using codoped Cu/N-TiO2 [205] and textile wastewaters
which was degraded more slowly than the parent molecule to be using Ti/TiO2 thin film decorated or not with Pt [204] and Ti/TiO2
completely mineralized. NTs [206]. Relevant results of some of these works are listed in
The addition of scavengers to the solution has demonstrated Table 7. These data encourage for a further intense research to show
the generation of oxidizing species in PEC. Fig. 26a exemplifies the benefits of PEC for treating many real wastewaters in order to
the abatement of the normalized sulfamethoxazole concentration develop a powerful technology at industrial level.
when 50 mL of 2.4 mg L−1 of this antibiotic in 10 mM NaCl of pH 2.7 Operation variables were optimized in some cases. For example,
Fang et al. [203] degraded 500 mL of a pharmaceutical wastewa-
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 29

H3C CH3 H3C CH3


OH N Cl OH N
CH3
OH OH

O O
OH OH
OH O OH O NH2 OH O OH O NH2
2 3

H3C CH3 H3C


Cl OH NH2 Cl OH N Cl OH NH
CH3 CH3 CH3
OH OH OH

O O O
OH OH OH
OH O OH O NH2 OH O OH O NH2 OH O OH O NH2
5 1 4

H3C CH3
Cl H3C N
OH H3C CH3
Cl H3C N
O OH
O
O
OH O
OH O OH O NH2
HO OH
6
OH O O O NH2
7
H3C CH3
Cl H3C N Cl H3C NH2
O OH
O O O
O
HO OH NH2 HO OH
OH H3C
OH O O O Cl H3C NH OH O O O NH2
8 OH 9
O
O
HO OH
OH O O O NH2
10
Fig. 25. Proposed reaction sequence for chlortetracycline degradation by PEC process using the arrangement of Fig. 11a with TiO2 thin-film photocatalyst upon 6.9 mW cm−2
UVC irradiation at 0.39 A. Adapted with permission from ref. [37].

Table 7
Per cent of color and COD removals for selected PEC treatments of real wastewaters using undivided reactors with different photoanodes.

Wastewater COD0 (mg L−1 ) Experimental conditions % color removal % COD decay Ref.

Pharmaceutical 3150 500 mL, 0.5 M NaCl, pH 3.0, Ni/TiO2 78 93 [203]


photocatalyst, Ecell = 10 V, 250 W UVA for 2 h
Landfill leachate 4378 500 mL, pH 2, 30 ◦ C, Cu/N-TiO2 photocatalyst, −a 69 [205]
Ecell = 20 V, 50 W visible light for 210 min
Textile effluent 153 8.5 L, pH 3, 2.9 g L−1 O3 , cylindrical Ti/TiO2 NTs 98 33 [206]
photocatalyst, Ecell = 2.0 V, 100 W UVB for
60 min
a
Not determined.

ter with COD0 = 3150 mg L−1 of pH 5.8 in an undivided cell similar irradiated on the photocatalyst in PC and PEC. The best PEC condi-
to that of Fig. 9a equipped with a photocatalyst composed of a tions were obtained by applying an Ecell = 10 V and operating with
Ni foam coated with TiO2 thin film and a cathode formed by 0.50 M NaCl and pH 3.0, showing the positive influence of the active
MnO2 and multi-walled carbon nanotube composite. Air injection chlorine species produced. Under these conditions, a loss of COD
ensured additional • OH generation from MnO2 reduction at the of 93%, much greater than 79% of color, was found for PEC (see
cathode during the EO and PEC processes and 250 W UVA was Table 7), values much higher than 66% and 53% for EO and 43% and
30 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

1.2
a
1.0

0.8
0
c/c

0.6

0.4

0.2

0.0
0 10 20 30 40 50 60 70 80 90
Time (min)
0.025 0.50
b

Photocurrent density (mA cm )


0.45
0.020
0.40

0.35
(min )

0.015
-1

0.30
dapp

0.010 0.25
k

0.20
0.005
0.15
-2

0.000 0.10
1 2 3
Treatment

Fig. 26. (a) Relative sulfamethoxazole concentration decay with time for the PEC
treatment of 50 mL of several antibiotic solutions in 10 mM NaCl of pH 2.7 using
an undivided cell with Ti/TiO2 thin-film photocatalyst at Eanod = +0.5 V (Ag/AgCl) Fig. 27. (a) Scheme of the annular bubble reactor developed for the degradation
under 4 W UVA light. (b) Apparent rate constant and photocurrent density for the of a real textile wastewater by direct ozonation (DO), DP, PC and PEC processes.
same trials. Treatment: ( ,1) 2.4 mg L−1 sulfamethoxazole, ( ,2) 2.4 mg L−1 sul- 1: reactor glass wall; 2: gas inlet; 3: sintered glass diffusers for injecting O2 /O3 ;
famethoxazole + 10 mg L−1 humic acid and ( ,3) 2.4 mg L−1 sulfamethoxazole + 10% 4: rubber tip; 5: glass tube to accommodate the UVB 100 W lamp; F: cylindrical
methanol. Adapted with permission from ref. [197]. TiO2 NTs; G: DSA used as cathode. (b) Electrical energy per order after 60 min of
the different treatments at pH 3 and 8 by injecting O3 at 2.9 g h−1 . Adapted with
permission from ref. [206].

59% for PC, demonstrating again the superiority of PEC over indi-
vidual processes. For a landfill leachate with COD0 = 4378 mg L−1 ,
ate ions, despite the production of • OH and other ROS from O3
TOC0 = 2583 mg L−1 and a very low BOD5 /COD = 0.01 of pH 7.8, Zhou
decomposition in alkaline medium:
et al. [205] studied the degradation by PEC of 500 mL wastewaters
in a cell like of Fig. 9a with a codoped Cu/N-TiO2 photoanode and O3 + OH− → • OH + (O2 •− ↔ HO2 • ) (32)
a graphite cathode, but without air injection. An outside 50 W vis-
For DO and coupled DP/O3 , PC/O3 and PEC/O3 trials, a compara-
ible light was utilized to illuminate the catalyst and pH 2, 30 ◦ C
tive energetic study was made from the electrical energy per order
and Ecell = 20 V were found as optimum operation variables, lead-
of magnitude consumed during 60 min, defined as follows:
ing to the best COD reduction of 69% in 210 min (see Table 7).
These authors also proposed a doubtful dynamic kinetic model to Electrical energy per order (kWh m−3 order−1 )
explain the COD decay rate as a function of COD, Ecell and pH, from
1000P t
which an activation energy of 63.5 kJ mol−1 for the process was = (33)
Vs log (color0 /color f )
estimated.
A more recent paper of Cardoso et al. [206] compared the degra- where P stands for the rated power (in kW) of the oxidation
dation of a real textile wastewater with color0 = 1080 mg Pt/Co L−1 system, t is the treatment time (in h), Vs denotes the treated vol-
and COD0 = 153 mg L−1 by direct ozonation (DO), DP, PC and PEC. ume (in L) and color0 and colorf mean the initial and final color
Additionally, the latter three processes were combined with ozone. measurements in Pt/Co units. As expected, Fig. 27b makes in evi-
Fig. 27a presents the 4 annular reactors used by bubbling O2 or O3 . dence an enhancement of this energetic parameter in the order:
The solution volume was 12 L for DO and 8.5 L for the other three DO < DP/O3 < PC/O3 < PEC/O3 , with greater values at pH 8 than at
photoassisted treatments due to the volume occupied by the inside pH 3. These differences may be cancelled by using free sunlight
100 W UVB lamp with ␭max = 315 nm. The PEC process was per- and solar photovoltaic panels, for example, to power the electrical
®
formed with a Ti/TiO2 NTs photoanode and a DSA cathode with devices, which need to be taken into account in the next future to
an optimum Ecell = 2.0 V. Poor color removal of 32%, 42% and 55%, design competitive PEC systems for wastewater remediation.
with scarce COD removal, was obtained by DP, PC and PEC by bub-
bling O2 at acidic pH 3 for 60 min, whereas 98% color decay with 5. Conclusions and prospects
33% COD abatement was found for DO at 2.9 g h−1 O3 flow rate.
When ozonation was coupled to the photoassisted processes, sim- Over the last 15 years, the potential applicability of PEC as
ilar results were always obtained. A slightly slower performance an emerging EAOP to the remediation of wastewaters containing
was obtained at pH 8 due to the presence of scavengers bicarbon- toxic and/or biorefractory organic pollutants has been well-proven.
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 31

It appears as one of the most highly promising water treatment Acknowlegments


technologies to deal with recalcitrant organic contaminants pol-
lution in water streams. The development of this technology has This research did not receive any specific grant from funding
been feasible thanks to the preparation of many novel semiconduc- agencies in the public, commercial, or not-for-profit sectors.
tors such as nanostructured, doped and composite TiO2 materials
with ability enough to absorb more light intensity and/or expand References
their photoactivity to the visible region compared with UV lamps
used to illuminate classical TiO2 thin-film, WO3 and ZnO photoan- [1] H. Zollinger, Color Chemistry: Syntheses, Properties, and Applications of
odes, initially checked for an effective separation of photogenerated Organic Dyes and Pigments, 3rd ed., VHCA and Wiley-VCH, Zurich, 2003.
[2] K.P. Sharma, S. Sharma, S. Sharma, P.K. Singh, S. Kumar, R. Grover, P.K.
electron/hole pairs. PEC systems have been designed at lab-scale, Sharma, A comparative study on characterization of textile wastewaters
usually being composed of tank reactors or flow cells with two- or (untreated and treated) toxicity by chemical and biological tests,
three-electrodes, upon illumination of the photoanode with inter- Chemosphere 69 (2007) 48–54.
[3] K. Kümmerer, The presence of pharmaceuticals in the environment due to
nal or external UVA, UVC, visible and/or Xe lamps. Good removal of human use −Present knowledge and future challenges, J. Environ. Manage.
color or initial substrate with smaller COD or TOC abatement were 90 (2009) 2354–2366.
generally obtained for synthetic wastewaters containing dyes, typ- [4] C.A. Damalas, I.G. Eleftherohorinos, Pesticide exposure safety issues, and risk
assessment indicators, Int. J. Environ. Res. Public Health 8 (2011) 1402–1419.
ical industrial chemicals and pharmaceuticals. Nevertheless, low [5] E. Brillas, C.A. Martinez-Huitle, Decontamination of wastewaters containing
organic loads with long treatment times were tested because of synthetic organic dyes by electrochemical methods. An updated review,
the low generation of oxidizing ROS due to the small photocur- Appl. Catal. B: Environ. 166–167 (2015) 603–643.
[6] M.A. Oturan, J.J. Aaron, Advanced oxidation processes in water/wastewater
rent produced. The superiority of PEC over individual processes was
treatment: principles and applications. A review, Crit. Rev. Environ. Sci.
well-demonstrated in many cases. The effect of operation variables Technol. 44 (2014) 2577–2641.
like pH, applied Eanod or janod , kind and concentration of electrolyte, [7] W.H. Glaze, J.-W. Kang, D.H. Chapin, The chemistry of water treatment
processes involving ozone, hydrogen peroxide and ultraviolet radiation,
pollutant concentration and temperature on the degradation per-
Ozone Sci. Eng. 9 (1987) 335–352.
formance of synthetic effluents was also analyzed to ascertain the [8] G.V. Buxton, C.L. Greenstock, W.P. Helman, A.B. Ross, Critical review of rate
optimum PEC conditions. The decay of low pollutant concentrations constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl
usually obeyed a pseudo-first-order kinetics, but high contents fol- radicals (OH/O− ) in aqueous solution, J. Phys. Chem. Ref. Data 17 (1988)
513–886.
lowed an adsorption Langmuir-Hinshelwood model in some cases. [9] E.G. Janzen, Y. Kotake, R.D. Hinton, Stabilities of hydroxyl radical spin
Scarce studies have been devoted to the detection of oxidation adducts of PBN-type spin traps, Free Radical Biol. Med. 12 (1992) 169–173.
products formed, the changes of toxicity and biodegradability, and [10] M.N. Chong, A.K. Sharma, S. Burn, C.P. Saint, Feasibility study on the
application of advanced oxidation technologies for decentralised
the estimation of the energy costs of the PEC process. An enhance- wastewater treatment, J. Clean. Prod. 35 (2012) 230–238.
ment of the oxidation ability of PEC was found for coupled PEC/O3 , [11] I. Sirés, E. Brillas, M.A. Oturan, M.A. Rodrigo, M. Panizza, Electrochemical
PEC/EF and PEF/SPEF systems. Promising good loses of color and advanced oxidation processes: today and tomorrow. A review, Environ. Sci.
Pollut. Res. 21 (2014) 8336–8367.
COD have been described for few real wastewaters by PEC and [12] B.P. Chaplin, Critical review of electrochemical advanced oxidation
PEC/O3 processes. processes for water treatment applications, Environ. Sci. Processes Impacts
On the basis of the present available data, more fundamental 16 (2014) 1182–1203.
[13] H.J. Lewerenz, C. Heine, K. Skorupska, N. Szabo, T. Hannappel, T. Vo-Dinh,
research is still needed to develop a mature PEC technology with
S.A. Campbell, H.W. Klemm, A.G. Munoz, Photoelectrocatalysis: principles,
perspectives to treating real wastewaters at industrial scale in the nanoemitter applications and routes to bio-inspired system, Energy Environ.
next future. It seems necessary to continue searching novel photo- Sci. 3 (2010) 748–760.
[14] A.G. Muñoz, H.J. Lewerenz, Advances in photoelectrocatalysis with
catalysts with enhanced photoactivity under sunlight irradiation to
nanotopographical photoelectrodes, ChemPhysChem 11 (2010) 1603–1615.
improve the actual photoefficiencies and PEC performance on pol- [15] V. Augugliaro, G. Camera-Roda, V. Loddo, G. Palmisano, L. Palmisano, J. Soria,
lutants abatement. These materials should be coupled to suitable S. Yurdakal, Heterogeneous photocatalysis and photoelectrocatalysis: from
reactors in pilot plants to show the viability of PEC degradation over unselective abatement of noxious species to selective production of
high-value chemicals, J. Phys. Chem. Lett. 6 (2015) 1968–1981.
real wastewaters for its further scale-up to industrial level. The effi- [16] T. Ochiai, A. Fujishima, Photoelectrochemical properties of TiO2
cient design of novel photoelectrochemical reactors will become a photocatalyst and its applications for environmental purification, J. Photoch.
future challenge to the scientist and engineers, since these reactors Photobio. C: Photochem. Rev. 13 (2012) 247–262.
[17] Y. Zhang, X. Xiong, Y. Han, X. Zhang, F. Shen, S. Deng, H. Xiao, X. Yang, G.
should embody the photocatalytical requirements in electrochem- Yang, H. Peng, Photoelectrocatalytic degradation of recalcitrant organic
ical reactors. In these systems, the use of free sunlight to illuminate pollutants using TiO2 film electrodes: an overview, Chemosphere 88 (2012)
the photoanodes, along with renewable energies such as photo- 145–154.
[18] R. Daghrir, P. Drogui, D. Robert, Photoelectrocatalytic technologies for
voltaic or wind devices to power the electrical instruments, should environmental applications, J. Photoch. Photobio. A: Chem. 238 (2012)
be designed to have a competitive process with minimum cost 41–52.
requirements for wastewater remediation. This point is highly fea- [19] A. Fujishima, X. Zhang, D.A. Tryk, TiO2 photocatalysis and related surface
phenomena, Surf. Sci. Rep. 63 (2008) 515–582.
sible due to the lower energy requirement of PEC systems as far
[20] D. Li, J. Qu, The progress of catalytic technologies in water purification: a
as we know. Since no overall removal of the organic pollutants is review, J. Environ. Sci. 21 (2009) 713–719.
expected, viability trials with pilot plants should include at least: (i) [21] I. Sirés, E. Brillas, Remediation of water pollution caused by pharmaceutical
residues based on electrochemical separation and degradation technologies
the optimization of color reduction and COD and/or TOC removal of
− A review, Environ. Int. 40 (2012) 212–229.
the treated effluent and (ii) the assessment of the decay in toxicity [22] J. Georgieva, E. Valova, S. Armyanov, N. Philippidis, I. Poulios, S. Sotiropoulos,
and the enhancement of biodregradability during the degradation Bi-component semiconductor oxide photoanodes for the
process. Coupling or combination with other powerful oxidation photoelectrocatalytic oxidation of organic solutes and vapors: a short
review with emphasis to TiO2 -WO3 photoanodes, J. Hazard. Mater. 211–212
techniques could also be investigated in order to develop treat- (2012) 30–46.
ments with higher oxidation power of the organic load. It can then [23] G.G. Bessegato, T.T. Guaraldo, J. Ferreira de Brito, M.F. Brugnera, M.V.B.
be envisaged that if the biodegradability of a given industrial efflu- Zanoni, Achievements and trends in photoelectrocatalysis: from
environmental to energy applications, Electrocatalysis 6 (2015) 415–441.
ent can be enhanced up to values >0.4 with an adequate single [24] X. Meng, Z. Zhang, X. Li, Synergetic photoelectrocatalytic reactors for
or coupled PEC process, the resulting wastewater could be post- environmental remediation: a review, J. Photochem. Photobiol. C:
treated by a biological process, suitable for high volumes of water, Photochem. Rev. 24 (2015) 83–101.
[25] M.A. Sousa, C. Gonçalves, V.J.P. Vilar, R.A.R. Boaventura, M.F. Alpendurada,
before disposal to the public sewage or reuse in human activities Suspended TiO2 -assisted photocatalytic degradation of emerging
such as agriculture or industrial manufacture of goods. contaminants in a municipal WWTP effluent using a solar pilot plant with
CPCS, Chem. Eng. J. 198–199 (2012) 301–309.
32 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

[26] L. Chen, C. Zhao, D. Dionysiou, K.E. O’shea, TiO2 photocatalytic degradation [54] L. Schmidt-Mende, J.L. MacManus-Driscoll, ZnO − nanostructures defects,
and detoxification of cylindrospermopsin, J. Photochem. Photobiol. A: Chem. and devices, Mater. Today 10 (2007) 40–48.
307–308 (2015) 115–122. [55] Ü. Özgür, Y.I. Alivov, C. Liu, A. Teke, M.A. Reshchikov, S. Doğan, V. Avrutin, S.J.
[27] P. Fernández-Ibánez, M.I. Polo-López, S. Malato, S. Wadhwa, J.W.J. Hamilton, Cho, H. Morkoç, A comprehensive review of ZnO materials and devices, J.
P.S.M. Dunlop, R. D’sa, E. Magee, K. O’shea, D.D. Dionysiou, J.A. Byrne, Solar Appl. Phys. 98 (2005) 1–103.
photocatalytic disinfection of water using titanium dioxide graphene [56] C. Klingshirn, ZnO: Material, physics and applications, ChemPhysChem 8
composites, Chem. Eng. J. 261 (2015) 36–44. (2007) 782–803.
[28] T. Yazawa, F. Machida, N. Kubo, T. Jin, Photocatalytic activity of transparent [57] A. Janotti, C.G. Van de Walle, Fundamentals of zinc oxide as a
porous glass supported TiO2 , Ceram. Int. 35 (2008) 3321–3325. semiconductor, Rep. Prog. Phys. 72 (2009) 126501 (29pp).
[29] G. Palmisano, V. Loddo, H.H. El Nazer, S. Yurdakal, V. Augugliaro, R. Criminna, [58] Z. Li, F. Gao, W. Kang, Z. Chen, M. Wu, L. Wang, D. Pan, Layer-by-layer
M. Pagliaro, Graphite-supported TiO2 for 4-nitrophenol degradation in a growth of ultralong ZnO vertical wire arrays for enhanced
photoelectrocatalytic reactor, Chem. Eng. J. 155 (2009) 339–346. photoelectrocatalytic activity, Mater. Lett. 97 (2013) 52–55.
[30] F. Mazille, T. Schoettl, N. Klamerth, S. Malato, C. Pulgarin, Field solar [59] M. Mahadik, S. Shinde, V. Mohite, S. Kumbhar, K. Rajpure, A. Moholkar, J.
degradation of pesticides and emerging water contaminants mediated by Kim, C. Bhosale, Photoelectrocatalytic oxidation of rhodamine B with
polymer films containing titanium and iron oxide with synergistic sprayed ␣-Fe2 O3 photocatalyst, Mater. Express 3 (2013) 247–255.
heterogeneous photocatalytic activity at neutral pH, Water Res. 44 (2010) [60] M. Zhang, W. Pu, S. Pan, O.K. Okoth, C. Yang, J. Zhang, Photoelectrocatalytic
3029–3038. activity of liquid phase deposited ␣-Fe2 O3 films under visible light
[31] S. Garcia-Segura, S. Dosta, J.M. Guilemany, E. Brillas, Solar illumination, J. Alloy. Compd. 648 (2015) 719–725.
photoelectrocatalytic degradation of Acid Orange 7 azo dye using a highly [61] S.E. LeBlanc, H.S. Fogler, The role of conduction/valence bands and redox
stable TiO2 photoanode synthesized by atmospheric plasma spray, Appl. potential in accelerated mineral dissolution, AIChE J. 32 (1986) 1702–1709.
Catal. B: Environ. 132–133 (2013) 142–150. [62] S. Yu, M. Xi, K. Han, Z. Wang, W. Yang, H. Zhu, Preparation and
[32] A. Fujishima, K. Honda, Electrochemical photolysis of water at a photoelectrocatalytic properties of polyaniline/layered manganese oxide
semiconductor electrode, Nature 238 (1972) 37–38. self-assembled film, Thin Solid Films 519 (2010) 357–361.
[33] J. O’M. Bockris, S.U.M. Khan, O.J. Murphy, M. Szklarczyk,;1 Technical brief on [63] C.M. Fan, B. Hua, Y. Wang, Z.H. Liang, X.G. Hao, S.B. Liu, Y.P. Sun, Preparation
models of photoelectrocatalysis, Int. J. Hydrogen Energy 9 (1984) 243–244. of Ti/SnO2 -Sb2 O4 photoanode by electrodeposition and dip coating for PEC
[34] A.Q. Contractor, J. O’M. Bockris, Photoelectrocatalysis of oxygen evolution on oxidations, Desalination 249 (2009) 736–741.
n-TiO2 , Electrochim. Acta 32 (1987) 121–123. [64] Y. Wang, C. Fan, B. Hua, Z. Liang, Y. Sun, Photoelectrocatalytic activity of two
[35] K. Nakata, A. Fujishima, TiO2 photocatalysis: design and applications, J. antimony doped SnO2 films for oxidation of phenol pollutants, Trans.
Photochem. Photobiol. C: Photochem. Rev. 13 (2012) 169–189. Nonferreous Met. Soc. China 19 (2009) 778–783.
[36] T. Hirakawa, C. Koga, N. Negishi, K. Takeuchi, S. Matsuzawa, An approach to [65] J. Florencio, M.J. Pacheco, A. Lopes, L. Ciriaco, Application of Ti/Pt/␤-PbO2
elucidating photocatalytic reaction mechanisms by monitoring dissolved anodes in the degradation of DR80 azo dye, Port. Electrochim. Acta 31
oxygen: effect of H2 O2 on photocatalysis, Appl. Catal. B: Environ. 87 (2009) (2013) 257–264.
46–55. [66] M. Rodrigues da Silva, A.C. Lucilha, R. Afonso, L.H. Dall’Antonia, L.V.A. Scalvi,
[37] R. Daghrir, P. Drogui, M.A. El Khakani, Photoelectrocatalytic oxidation of Photoelectrochemical properties of FTO/m-BiVO4 electrode in different
chlortetracycline using Ti/TiO2 photo-anode with simultaneous H2 O2 electrolytes solutions under visible light irradiation, Ionics 20 (2014)
production, Electrochim. Acta 87 (2013) 18–31. 105–113.
[38] L.C. Almeida, B.F. Silva, M.V.B. Zanoni, [67] S. Sun, W. Wang, L. Zhang, Efficient contaminant removal by Bi2 WO6 films
Photoelectrocatalytic/photoelectro-Fenton coupling system using a with nanoleaflike structures through a photoelectrocatalytic process, J.
nanostructured photoanode for the oxidation of a textile dye: kinetics study Phys. Chem. C 116 (2012) 19413–19418.
and oxidation pathway, Chemosphere 136 (2015) 63–71. [68] C.S. Pan, Y.F. Zhu, New type of BiPO4 oxy-acid salt photocatalyst with high
[39] Y. Nosaka, A. Nosaka, Understanding hydroxyl radical (OH) Generation photocatalytic activity on degradation of dye, Environ. Sci. Technol. 44
Processes in Photocatalysis, ACS Energy Lett. 1 (2016) 356–359. (2010) 5570–5574.
[40] M. Panizza, G. Cerisola, Direct and mediated anodic oxidation of organic [69] W. Liao, J. Yang, H. Zhou, M. Murugananthan, Y. Zhang, Electrochemically
pollutants, Chem. Rev. 109 (2009) 6541–6569. self-doped TiO2 nanotube arrays for efficient visible light
[41] E.B. Cavalcanti, S. García-Segura, F. Centellas, E. Brillas, Electrochemical photoelectrocatalytic degradation of contaminants, Electrochim. Acta 136
incineration of omeprazole in neutral aqueous medium using a platinum or (2014) 310–317.
boron-doped diamond. Degradation kinetics and oxidation products, Water [70] Q. Chen, H. Liu, Y. Xin, X. Cheng, Coupling immobilized TiO2 nanobelts and
Res. 47 (2013) 1803–1815. Au nanoparticles for enhanced photocatalytic and photoelectrocatalytic
[42] X. Florenza, A.M.S. Solano, F. Centellas, C.A. Martínez-Huitle, E. Brillas, S. activity and mechanism insights, Chem. Eng. J. 241 (2014) 145–154.
Garcia-Segura, Degradation of the azo dye Acid Red 1 by anodic oxidation [71] J. Gong, W. Pu, C. Yang, J. Zhang, A simple electrochemical oxidation method
and indirect electrochemical processes based on Fenton’s reaction to prepare highly ordered Cr-doped titania nanotube arrays with promoted
chemistry. Relationship between decolorization, mineralization and photoelectrochemical property, Electrochim. Acta 68 (2012) 178–183.
products, Electrochim. Acta 142 (2014) 276–288. [72] O. Kerkez, I. Boz, Photo(electro)catalytic activity of Cu2+ -modified TiO2
[43] P.J. Huang, H. Chang, C.T. Yeh, C.W. Tsai, Phase transformation of TiO2 nanorod array thin films under visible light irradiation, J. Phys. Chem. Solids
monitored by Thermo-Raman spectroscopy with TGA/DTA, Termochim. 75 (2014) 611–618.
Acta 297 (1997) 85–92. [73] W. Tang, X. Chen, J. Xia, J. Gong, X. Zeng, Preparation of an Fe-doped
[44] W.H. Leng, Z. Zhang, J.Q. Zhang, Photoelectrocatalytic degradation of aniline visible-light-response TiO2 film electrode and its photoelectrocatalytic
over rutile TiO2 /Ti electrode thermally formed at 600 ◦ C, J. Mol. Catal. A: activity, Mater. Sci. Eng. B: Adv. 187 (2014) 39–45.
Chem. 206 (2003) 239–252. [74] G.C. Bessegato, J.C. Cardoso, M.V.B. Zanoni, Enhanced photoelectrocatalytic
[45] A. Di Paola, M. Bellardita, L. Palmisano, Brookite, the least known TiO2 degradation of an acid dye with boron-doped TiO2 nanotube anodes, Catal.
photocatalyst, Catalysts 3 (2013) 36–73. Today 240 (2015) 100–106.
[46] T. Luttrell, S. Halpegamage, J. Tao, A. Kramer, E. Sutter, M. Batzill, Why is [75] Z.-L. Cheng, S. Han, Preparation and photoelectrocatalytic performance of
anatase a better photocatalyst than rutile? −Model studies on epitaxial TiO2 N-doped TiO2 /NaY zeolite membrane composite electrode material, Water
films, Sci. Rep. 4 (2015) 4043–4050. Sci. Technol. 73 (2016) 486–492.
[47] R. Kaplan, B. Erjavec, G. Drazić, J. Grdadolnik, A. Pintar, Simple synthesis of [76] Z. Zhang, Y. Yu, P. Wang, Hierarchical top-porous/bottom-tubular TiO2
anatase/rutile/brookite TiO2 nanocomposite with superior mineralization nanostructures decorated with Pd nanoparticles for efficient
potential for photocatalytic degradation of water pollutants, Appl. Catal. B: photoelectrocatalytic decomposition of synergistic pollutants, ACS Appl.
Environ. 18 (2016) 465–474. Mater. Interfaces 4 (2012) 990–996.
[48] E. Lassner, W.D. Schubert, Tungsten: Properties, Chemistry, Technology of [77] K. Xie, L. Sun, C. Wang, Y. Lai, M. Wang, H. Chen, C. Lin, Photoelectrocatalytic
the Element, Alloys and Chemical Compounds, first ed., Springer, New York, properties of Ag nanoparticles loaded TiO2 nanotube arrays prepared by
1999. pulse current deposition, Electrochim. Acta 55 (2010) 7211–7218.
[49] M. Yagi, S. Maruyama, K. Sone, K. Nagai, T. Norimatsu, Preparation and [78] A. Pandikumar, S. Murugesan, R. Ramaraj, Functionalized silicate
photoelectrocatalytic activity of a nano-structured WO3 platelet film, J. Solid sol-gel-supported TiO2 -Au core-shell nanomaterials and their
State Chem. 181 (2008) 175–182. photoelectrocatalytic activity, ACS Appl. Mater. Interface 2 (2010)
[50] Z. Xu, X. Li, J. Li, L. Wu, Q. Zeng, Z. Zhou, Effect of CoOOH loading on the 1912–1917.
photoelectrocatalytic performance of WO3 nanorod array film, Appl. Surf. [79] Z. Zhou, L. Zhu, J. Li, H. Tang, Electrochemical preparation of TiO2 /SiO2
Sci. 284 (2013) 285–290. composite film and its high activity toward the photoelectrocatalytic
[51] J.M. Stellman, Encyclopaedia of Occupational Health and Safety: Chemical, degradation of methyl orange, J. Appl. Electrochem. 39 (2009) 1745–1753.
Industries and Occupations, fourth ed., International Labour Office, Geneva, [80] L.E. Fraga, J.H. Franco, M.O. Orlandi, M.V.B. Zanoni, Photoelectrocatalytic
1998. oxidation of hair dye basic red 51 at W/WO3 /TiO2 bicomposite photoanode
[52] M. Hepel, J. Luo, Photoelectrochemical mineralization of textile diazo dye activated by ultraviolet and visible radiation, J. Env. Chem. Eng. 1 (2013)
pollutants using nanocrystalline WO3 electrodes, Electrochim. Acta 47 194–199.
(2001) 729–740. [81] X. Hu, J. Yang, J. Zhang, Magnetic loading of TiO2 /SiO2 /Fe3 O4 nanoparticles
[53] J. Luo, M. Hepel, Photoelectrochemical degradation of naphthol blue black on electrode surface for photoelectrocatalytic degradation of diclofenac, J.
diazo dye on WO3 film electrode, Electrochim. Acta 46 (2001) 2913–2922. Hazard. Mater. 196 (2011) 220–227.
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 33

[82] S. Chai, G. Zhao, P. Li, Y. Lei, Y. Zhang, D. Li, Novel sieve-like SnO2 /TiO2 [109] X. Cheng, G. Pan, X. Yu, Construction of high-efficient photoelectrocatalytic
nanotubes with integrated photoelectrocatalysis: fabrication and system by coupling with TiO2 nano-tubes photoanode and active
application for efficient toxicity elimination of nitrophenol wastewater, J. carbono/polytetrafluoroethylene cathode and its enhanced
Phys. Chem. C 115 (2011) 18261–18269. photoelectrocatalytic degradation of 2,4-dichlorophene and mechanism,
[83] W. Siripala, A. Ivanovskaya, T.F. Jaramillo, S. Baeck, E.W. McFarland, A Chem. Eng. J. 279 (2015) 264–272.
Cu2 O/TiO2 heterojunction thin film cathode for photoelectrocatalysis, Sol. [110] S.K. Mohapatra, M. Misra, V.K. Mahajan, K.S. Raja, A novel method for the
Energ. Mat. Sol C 77 (2003) 229–237. synthesis of titania nanotubes using sonoelectrochemical method and its
[84] G. Li, L. Wu, F. Li, P. Xu, D. Zhang, H. Li, Photoelectrocatalytic degradation of application for photoelectrochemical splitting water, J. Catal. 246 (2007)
organic pollutants via a CdS quantum dots enhanced TiO2 nanotube array 362–369.
electrode under visible light irradiation, Nanoscale 5 (2013) 2118–2125. [111] Y. Liu, K. Mu, G. Yang, H. Peng, F. Shen, L. Wang, S. Deng, X. Zhang, Y. Zhang,
[85] G.G. Bessegato, J.C. Cardoso, B. Ferreira da Silva, M.V.B. Zanoni, Enhanced Fabrication of a coral/double-wall TiO2 nanotube array film electrode with
photoabsorption properties of composites of Ti/TiO2 nanotubes decorated higher photoelectrocatalytic activity under sunlight, New. J. Chem. 39
by Sb2 S3 and improvement of degradation of hair dye, J. Photochem. (2015) 3923–3928.
Photobiol. A: Chem. 276 (2013) 96–103. [112] S. Dosta, M. Robotti, S. Garcia-Segura, E. Brillas, I.G. Cano, J.M. Guilemany,
[86] L. Yang, Z. Li, H. Jiang, W. Jiang, R. Su, S. Luo, Y. Luo, Photoelectrocatalytic Influence of atmospheric plasma spraying on the solar
oxidation of bisphenol A over mesh of TiO2 /graphene/Cu2 O, Appl. Catal. B: photoelectron-catalytic properties of TiO2 coatings, Appl. Catal. B: Environ.
Environ. 183 (2016) 75–85. 189 (2016) 151–159.
[87] O. Zhang, B. Lin, Y. Chen, B. Gao, L. Fu, B. Li, Electrochemical and [113] J.F. Fu, Y.Q. Zhao, X.D. Xue, W.C. Li, A.O. Babatunde, Multivariate-parameter
photoelectrochemical characteristics of TiNbO5 nanosheet electrode, optimization of acid blue-7 wastewater treatment by Ti/TiO2
Electrochim. Acta 81 (2012) 74–82. photoelectrocatalysis via the Box-Behnken design, Desalination 243 (2009)
[88] H.S. Kushwaha, N.A. Madhar, B. Ilahi, P. Thomas, A. Halder, R. Vaish, Efficient 42–51.
solar energy conversion using CaCu3 Ti4 O12 photoanode for photocatalysis [114] Y. Su, S. Han, X. Zhang, X. Chen, L. Lei, Preparation and visible-light-driven
and photoelectrocatalysis, Sci. Reports 6 (2016) 18557. photoelectrocatalytic properties of boron-doped TiO2 nanotubes, Mater.
[89] J.C. Cardoso, T.M. Lizier, M.V.B. Zanoni, Highly ordered TiO2 nanotube arrays Chem. Phys. 110 (2008) 239–246.
and photoelectrocatalytic oxidation of aromatic amine, Appl. Catal. B: [115] M.J. Martin de Vidales, L. Mais, C. Saez, P. Canizares, F.C. Walsh, M.A.
Environ. 99 (2010) 96–102. Rodrigo, C.D.A. Rodrigues, C. Ponce de Leon, Photoelectrocatalytic oxidation
[90] M.V.B. Zanoni, J.J. Sene, M.A. Anderson, Photoelectrocatalytic degradation of of methyl orange on a TiO2 nanotubular anode using a flow cell, Chem. Eng.
Remazol Brilliant Orange 3R on titanium dioxide thin-film electrodes, J. Technol. 39 (2016) 135–141.
Photoch. Photobio. A: Chem. 157 (2003) 55–63. [116] H. Selcuk, J.J. Sene, M.A. Anderson, Photoelectrocatalytic humic acid
[91] W.Y. Gan, M.W. Lee, R. Amal, H. Zao, K. Chiang, Photoelectrocatalytic activity degradation kinetics and effect of pH, applied potential and inorganic ions, J.
of mesoporous TiO2 films prepared using the sol-gel method with tri-block Chem. Technol. Biotechnol. 78 (2003) 979–984.
copolymer as structure directing agent, J. Appl. Electrochem. 38 (2008) [117] X. Ding, Z. Ai, L.A. Zhang, Dual-cell wastewater treatment system with
703–712. combining anodic visible light driven photoelectro-catalytic oxidation and
[92] J. Marugán, P. Christensen, T. Egerton, H. Purnama, Synthesis, cathodic electro-Fenton oxidation, Sep. Purif. Technol. 125 (2014) 103–110.
characterization and activity of photocatalytic sol-gel TiO2 powders and [118] F.C. Moreira, R.A.R. Boaventura, E. Brillas, V.J.P. Vilar, Electrochemical
electrodes, Appl. Catal. B: Environ. 89 (2009) 273–283. advanced oxidation processes: a review on their application to synthetic
[93] J. Li, L. Li, L. Zheng, Y. Xian, L. Jin, Photoelectrocatalytic degradation of and real wastewaters, Appl. Catal. B: Environ. 202 (2017) 217–261.
rhodamine B using Ti/TiO2 electrode prepared by laser calcination method, [119] E.J. Ruiz, C. Arias, E. Brillas, A. Hernández-Ramírez, J.M. Peralta-Hernández,
Electrochim. Acta 51 (2006) 4942–4949. Mineralization of Acid Yellow 36 azo dye by electro-Fenton and solar
[94] P.S. Shinde, S.B. Sadale, P.S. Patil, P.N. Bhosale, A. Brüger, M. photoelectro-Fenton processes with a boron-doped diamond anode,
Neumann-Spallart, C.H. Bhosale, Properties of spray deposited titanium Chemosphere 82 (2011) 495–501.
dioxide thin films and their application in photoelectrocatalysis, Sol. Energ. [120] M.E. Olya, A. Pirkarami, M. Soleimani, M. Bahmaei, Photoelectrocatalytic
Mat. Sol. C 92 (2008) 283–290. degradation of acid dye using Ni-TiO2 with the energy supplied by solar cell:
[95] P.S. Shinde, P.S. Patil, P.N. Bhosale, A. Brüger, G. Nauer, M. mechanism and economical studies, J. Environ. Manage. 121 (2013)
Neumann-Spallart, C.H. Bhosale, UVA and solar light assisted 210–219.
photoelectrocatalytic degradation of AO7 dye in water using spray [121] A. Pirkarami, M.E. Olya, S.R. Farshid, UV/Ni-TiO2 nanocatalyst for
deposited TiO2 thin films, Appl. Catal. B: Environ. 89 (2009) 288–294. electrochemical removal of dyes considering operating costs, Water Res.
[96] L. Sirghi, T. Aoki, Y. Hatanaka, Friction force microscopy study of the Ind. 5 (2014) 9–20.
hydrophilicity of TiO2 thin films deposited by radio frequency magnetron [122] F.-B. Li, X.-Z. Li, Y.-H. Kang, X.-J. Li, An innovative Ti/TiO2 mesh
sputtering, Surf. Rev. Lett. 10 (2003) 345–349. photoelectrode for methyl orange photoelectrocatalytic degradation, J.
[97] C. He, X.Z. Li, N. Graham, Y. Wang, Preparation of TiO2 /ITO and TiO2 /Ti Environ. Sci. Health A: Toxic/Hazard. Subst. Environ. Eng. A37 (2002)
photoelectrodes by magnetron sputtering for photocatalytic application, 623–640.
Appl. Catal. A: Gen. 305 (2006) 54–63. [123] M. Li, L. Xiong, Y. Chen, N. Zhang, Y. Zhang, H. Yin, Studies on
[98] G. Li Puma, A. Bono, J.G. Collin, Preparation of titanium dioxide photo-electro-chemical catalytic degradation of Acid Scarlet 3R dye, Sci.
photocatalyst loaded onto activated carbon support using chemical vapour China B: Chem. 48 (2005) 297–304.
deposition: a review paper, J. Hazard. Mater. 157 (2008) 209–219. [124] Y.-F. Su, T.-C. Chou, Comparison of the photocatalytic and
[99] M.L. Hitchman, F. Tian, Studies of TiO2 thin films prepared by chemical photoelectrocatalytic decolorization of methyl orange on sputtered TiO2
vapour deposition for photocatalytic and photoelectrocatalytic degradation thin films, Z. Naturfor. B: Chem. Sci. 60 (2005) 1158–1167.
of 4-chlorophenol, J. Electroanal. Chem. 538-539 (2002) 165–172. [125] T.A. Egerton, H. Purnama, S. Purwajanti, M. Zafar, Decolourization of dye
[100] M. Heikkilä, E. Puukilainen, M. Ritala, M. Leskelä, Effect of thickness of ALD solutions using photoelectrocatalysis and photocatalysis, J. Adv.
grown TiO2 films on photoelectrocatalysis, J. Photochem. Photobiol. A: Oxid.Technol. 9 (2006) 79–85.
Chem. 204 (2009) 200–208. [126] Y.B. Xie, X.Z. Li, Interactive oxidation of photoelectrocatalysis and
[101] T. Wang, Z. Luo, C. Li, J. Gong, Controllable fabrication of nanostructured electro-Fenton for azo dye degradation using TiO2 -Ti mesh and reticulated
materials for photoelectrochemical water splitting via atomic layer vitreous carbon electrodes, Mater. Chem. Phys. 95 (2006) 39–50.
deposition, Chem. Soc. Rev. 43 (2014) 7469–7484. [127] L.C. Macedo, D.A.M. Zaia, G.J. Moore, H. de Santana, Degradation of leather
[102] M. Gardon, J.M. Guilemany, Milestones in functional titanium dioxide dye on TiO2 : a study of applied experimental parameters on
thermal spray coatings: a review, J. Therm. Spray Technol. 23 (2014) photoelectrocatalysis, J. Photochem. Photobiol. A: Chem. 185 (2007) 86–93.
577–595. [128] B. Liu, L. Wen, X. Zhao, Efficient degradation of aqueous methyl orange over
[103] Y. Su, X. Zhang, M. Zhou, S. Han, L. Lei, Preparation of high efficient TiO2 and CdS electrodes using photoelectrocatlysis under UV and visible
photoelectrode of N-F-codoped TiO2 nanotubes, J. Photochem. Photobiol. A: light irradiation, Progress Org. Coat. 64 (2009) 120–123.
Chem. 194 (2008) 152–160. [129] J. Li, J. Wang, L. Huang, G. Lu, Photoelectrocatalytic degradation of methyl
[104] Y. Xie, D. Fu, Photoelectrocatalysis reactivity of independent titania orange over mesoporous film electrodes, Photochem. Photobiol. Sci. 9
nanotubes, J. Appl. Electrochem. 40 (2010) 1281–1291. (2010) 39–46.
[105] Q. Wang, J. Shang, T. Zhu, F. Zhao, Efficient photoelectrocatalytic reduction [130] M. Neumann-Spallart, Photoelectrochemistry on a planar, interdigitated
of Cr(VI) using TiO2 nanotube arrays as the photoanode and a large-area electrochemical cell, Electrochim. Acta 56 (2011) 8752–8757.
titanium mesh as the photocathode, J. Mol. Catal. A: Chem. 335 (2011) [131] H. Su, H. Du, Study on photoelectrocatalysis of three-dimensional electrode
242–247. using TiO2 coatings particle electrode, Adv. Mater. Res. 156–157 (2011)
[106] A. Zhang, M. Zhou, L. Liu, W. Wang, Y. Jiao, Q. Zhou, A novel 344–349.
photoelectrocatalytic system for organic contaminant degradation on a TiO2 [132] W. Zhanga, J. Bai, J. Fu, Photoelectrocatalytic degradation of methyl orange
nanotube (TNT)/Ti electrode, Electrochim. Acta 55 (2010) 5091–5099. on porous TiO2 film electrode in NaCl solution, Adv. Mater. Res. 213 (2011)
[107] H. Liu, X. Cao, G. Liu, Y. Wang, N. Zhang, T. Li, R. Tough, Photoelectrocatalytic 15–19.
degradation of triclosan on TiO2 nanotube arrays and toxicity change, [133] W. Zhang, Q. Li, H. He, Photoelectrocatalytic properties of porous TiO2 films
Chemosphere 93 (2013) 160–165. prepared using ODA as template, Adv. Mater. Res. 457–458 (2012) 521–524.
[108] D. Tekin, B. Saygi, Photoelectrocatalytic decomposition of acid black 1 dye [134] T.N.M. Cervantes, D.A.M. Zaia, G.J. Moore, H. Santana, Photoelectrocatalysis
using TiO2 nanotubes, J. Environ. Chem. Eng. 1 (2013) 1057–1061. study of the decolorization of synthetic azo dye mixtures on Ti/TiO2 ,
Electrocatalysis 4 (2013) 85–91.
34 S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35

[135] B. Tan, Y. Zhang, M. Long, Large-scale preparation of nanoporous TiO2 film [160] L.E. Fraga, M.V.B. Zanoni, Nanoporous of W/WO3 thin film electrode grown
on titanium substrate with improved photoelectrochemical performance, by electrochemical anodization applied in the photoelectrocatalytic
Nanoscale Res. Lett. 9 (2014) 190 (6 p.). oxidation of the Basic Red 51 used in hair dye, J. Braz. Chem. Soc. 22 (2011)
[136] C.-F. Liu, C.P. Huang, C.-C. Hu, Y. Juang, C. Huang, Photoelectrochemical 718–725.
degradation of dye wastewater on TiO2 -coated titanium electrode prepared [161] P. Chatchai, A.Y. Nosaka, Y. Nosaka, Photoelectrocatalytic performance of
by electrophoretic deposition, Sep. Purif. Technol. 165 (2016) 145–153. WO3 /BiVO4 toward the dye degradation, Electrochim. Acta 94 (2013)
[137] L.G. Devi, N. Kottam, Enhanced activity of nano-TiO2 in photoelectrocatalysis 314–319.
over photocatalysis process for the degradation of Indanthrene BR Violet [162] S.S. Shinde, C.H. Bhosale, K.Y. Rajpure, Solar light assisted photocatalysis of
dye in presence of UV-light, Polish J. Chem. 81 (2007) 1819–1827. water using a zinc oxide semiconductor, J. Semicond. 34 (2013),
[138] M.E. Osugi, M.V.B. Zanoni, C.R. Chenthamarakshan, N.R. de Tacconi, G.A. 043002/1-043002/4.
Woldemariam, S.S. Mandal, K. Rajeshwar, Toxicity assessment and [163] Q. Zeng, J. Bai, J. Li, L. Xia, K. Huang, X. Li, B. Zhou, A novel in situ preparation
degradation of disperse azo dyes by photoelectrocatalytic oxidation on method for nanostructured ␣-Fe2 O3 films from electrodeposited Fe films for
Ti/TiO2 nanotubular array electrodes, J. Adv. Oxid. Technol. 11 (2008) efficient photoelectrocatalytic water splitting and the degradation of organic
425–434. pollutants, J. Mater. Chem. A: Mater. Energy Sustain. 3 (2015) 4345–4353.
[139] J. Bai, B. Zhou, J. Li, W. Cai, Enhanced photoelectrocatalytic degradation of [164] T. Zeng, X. Yu, K.-H. Ye, Z. Qiu, Y. Zhu, Y. Zhang, BiPO4 film on ITO substrates
azo-dye pollutants using transparent titania nanotube arrays glass for photoelectrocatalytic degradation, Inorg. Chem. Commun. 58 (2015)
electrode, Adv. Mater. Res. 311–313 (2011) 2089–2092. 39–42.
[140] E.R.A. Ferraz, G.A.R. Oliveira, M.D. Grando, T.M. Lizier, M.V.B. Zanoni, D.P. [165] Y. Cong, J. Wang, H. Jin, X. Feng, Q. Wang, Y. Ji, Y. Zhang, Enhanced
Oliveira, Photoelectrocatalysis based on Ti/TiO2 nanotubes removes toxic photoelectrocatalytic activity of a novel Bi2 O3 -BiPO4 composite electrode
properties of the azo dyes Disperse Red 1, Disperse Red 13 and Disperse for the degradation of refractory pollutants under visible light irradiation,
Orange 1 from aqueous chloride samples, J. Environ. Manage. 124 (2013) Ind. Eng. Chem. Res. 55 (2016) 1221–1228.
108–114. [166] S.V. Mohite, V.V. Ganbavle, K.Y. Rajpure, Solar photoelectrocatalytic
[141] M.V.B. Zanoni, T.T. Guaraldo, Photoelectrochemical hydrogen generation activities of rhodamine-B using sprayed WO3 photoelectrode, J. Alloys
and concomitant organic dye oxidation under TiO2 nanotube, ECS Trans. 50 Comp. 655 (2016) 106–113.
(2013) 63–70. [167] S.M.A. Jorge, J.J. de Sene, A.O. Florentino, Photoelectrocatalytic treatment of
[142] S. Franz, D. Perego, O. Marchese, M. Bestetti, Photoelectrochemical advanced p-nitrophenol using Ti/TiO2 thin-film electrode, J. Photochem. Photobiol. A:
oxidation processes on nanostructured TiO2 catalysts: decolorization of a Chem. 174 (2005) 71–75.
textile azo-dye, Khim. Tekhnol. Vody 37 (2015) 207–219. [168] T. An, Y. Xiong, G. Li, X. Zhu, G. Sheng, J. Fu, Improving ultraviolet light
[143] G.G. Bessegato, J.C. Cardoso, B. Ferreira da Silva, M.V.B. Zanoni, Combination transmission in a packed-bed photoelectrocatalytic reactor for removal of
of photoelectrocatalysis and ozonation: a novel and powerful approach oxalic acid from wastewater, J. Photochem. Photobiol. A: Chem. 181 (2006)
applied in Acid Yellow 1 mineralization, Appl. Catal. B: Environ. 180 (2016) 158–165.
161–168. [169] H. Selcuk, M. Bekbolet, Photocatalytic and photoelectrocatalytic humic acid
[144] E.C. Rodrigues, L.A. Soares, J.A. Modenes Jr., J.J. Sene, G. Bannach, C. Carvalho, removal and selectivity of TiO2 coated photoanode, Chemosphere 73 (2008)
M. Ionashiro, Synthesis and characterization of Fe(III)-doped ceramic 854–858.
membranes of titanium dioxide and its application in photoelectrocatalysis [170] L.E. Fraga, M.A. Anderson, M.L.P.M.A. Beatriz, F.M.M. Paschoal, L.P. Romao,
of a textile dye, Eclet. Quim. 36 (2011) 0002, 19 pp. M.V.B. Zanoni, Evaluation of the photoelectrocatalytic method for oxidizing
[145] J. Gong, W. Pu, C. Yang, J. Zhang, Tungsten and nitrogen co-doped TiO2 chloride and simultaneous removal of microcystin toxins in surface waters,
electrode sensitized with Fe-chlorophyllin for visible light Electrochim. Acta 54 (2009) 2069–2076.
photoelectrocatalysis, Chem. Eng. J. 209 (2012) 94–101. [171] H. Chen, D. Li, X. Li, J. Li, Q. Chen, B. Zhou, Adsorption and
[146] H. Liao, W. Zhang, X. Sun, L. Shi, M. Qin, Synthesis and photoelectrocatalytic photoelectrocatalytic characteristics of organics on TiO2 nanotube arrays, J.
property of two-nonmetal-codoped TiO2 nanotube arrays with high aspect Solid State Electrochem. 16 (2012) 3907–3914.
ratio, Adv. Mater. Res. 412 (2012) 219–222. [172] V.M. Daskalaki, I. Fulgione, Z. Frontistis, L. Rizzo, D. Mantzavinos, Solar
[147] F.-J. Zhang, M.-L. Chen, W.-C. Oh, Visible light photoelectrocatalytic light-induced photoelectrocatalytic degradation of bisphenol-A on TiO2 /ITO
degradation of methylene blue by platinum-treated carbon film anode and BDD cathode, Catal. Today 209 (2013) 74–78.
nanotube/titania composites, Asian J. Chem. 22 (2010) 5636–5648. [173] L. Suhadolnik, A. Pohar, B. Likozar, M. Ceh, Mechanism and kinetics of phenol
[148] G. Wang, S. Chen, H. Yu, X. Quan, Integration of membrane filtration and photocatalytic, electrocatalytic and photoelectrocatalytic degradation in a
photoelectrocatalysis using a TiO2 /carbon/Al2 O3 membrane for enhanced TiO2 -nanotube fixed-bed microreactor, Chem. Eng. J. 303 (2016) 292–301.
water treatment, J. Hazard. Mater. 299 (2015) 27–34. [174] J. Gong, C. Yang, W. Pu, J. Zhang, Liquid phase deposition of tungsten doped
[149] Y.-N. Zhang, G. Zhao, Y. Lei, P. Li, M. Li, Y. Jin, B. Lv, CdS-encapsulated TiO2 TiO2 films for visible light photoelectrocatalytic degradation of
nanotube arrays lidded with ZnO nanorod layers and their dodecyl-benzenesulfonate, Chem. Eng. J. 167 (2011) 190–197.
photoelectrocatalytic applications, ChemPhysChem 11 (2010) 3491–3498. [175] K. Yang, W. Pu, Y. Tan, M. Zhang, C. Yang, J. Zhang, Enhanced
[150] Y. Shen, F. Li, D. Liu, S. Li, L. Fan, Fabrication of TiO2 nanotube films modified photoelectrocatalytic activity of Cr-doped TiO2 nanotubes modified with
with Ag2 S and photoelectrocatalytic decolorization of methyl orange under polyaniline, Mater. Sci. Semicond. Process. 27 (2014) 777–784.
solar light, Sci. Adv. Mater. 4 (2012) 1214–1219. [176] C. He, Y. Xiong, J. Chen, C. Zha, X. Zhu, Photoelectrochemical performance of
[151] D. Wang, X. Li, J. Chen, X. Tao, Enhanced photoelectrocatalytic activity of Ag-TiO2 /ITO film and photoelectrocatalytic activity towards the oxidation of
reduced graphene oxide/TiO2 composite films for dye degradation, Chem. organic pollutants, J. Photochem. Photobiol. A: Chem. 157 (2003) 71–79.
Eng. J. 198–199 (2012) 547–554. [177] X.Z. Li, C. He, N. Graham, Y. Xiong, Photoelectrocatalytic degradation of
[152] M. Li, G. Zhao, P. Li, Y. Zhang, M. Wu, Photoelectrocatalytic properties of a bisphenol A in aqueous solution using Au-TiO2 /ITO film, J. Appl.
vertically aligned Ti-W alloy oxide nanotubes array and its applications in Electrochem. 35 (2005) 741–750.
dye wastewater degradation, Environ. Technol. 33 (2012) 191–199. [178] M. Antoniadou, V.M. Daskalaki, N. Balis, D.I. Kondarides, C. Kordulis, P.
[153] Q. Wang, J. Shang, H. Song, T. Zhu, J. Ye, F. Zhao, J. Li, S. He, Visible-light Lianos, Photocatalysis and photoelectrocatalysis using (CdS-ZnS)-TiO2
photoelectrocatalytic degradation of rhodamine B over planar devices using combined photocatalysts, Appl. Catal. B: Environ. 107 (2011) 188–196.
a multi-walled carbon nanotube-TiO2 composite, Mater. Sci. Semicond. [179] H. Su, H. Du, Study on photoelectrocatalytic of three-dimensional electrode
Process. 16 (2013) 480–484. using TiO2 coated ␥-Al2 O3 and scrap iron particle electrode, Appl. Mech.
[154] C. Zhai, M. Zhu, F. Ren, Z. Yao, Y. Du, P. Yang, Enhanced photoelectrocatalytic Mater. 71–78 (2011) 972–975.
performance of titanium dioxide/carbon cloth based photoelectrodes by [180] H. Su, H. Du, Photoelectrocatalytic performance of TiO2 film treatment of
graphene modification under visible-light irradiation, J. Hazard. Mater. 263 La(NO3 )3 , Energy Procedia 11 (2011) 2333–2338.
(2013) 291–298. [181] X. Zhao, H. Liu, J. Qu, Photoelectrocatalytic degradation of organic
[155] M. Zhang, C. Yang, W. Pu, Y. Tan, K. Yang, J. Zhang, Liquid phase deposition of contaminants at Bi2 O3 /TiO2 nanotube array electrode, Appl. Surf. Sci. 257
WO3 /TiO2 heterojunction films with high photoelectrocatalytic activity (2011) 4621–4624.
under visible light irradiation, Electrochim. Acta 148 (2014) 180–186. [182] W. Liao, Y. Zhang, M. Zhang, M. Murugananthan, S. Yoshihara,
[156] T.T. Guaraldo, V.R. Goncales, B.F. Silva, S.I.C. de Torresi, M.V.B. Zanoni, Photoelectrocatalytic degradation of microcystin-LR using Ag/AgCl/TiO2
Hydrogen production and simultaneous photoelectrocatalytic pollutant nanotube arrays electrode under visible light irradiation, Chem. Eng. J. 231
oxidation using a TiO2 /WO3 nanostructured photoanode under visible light (2013) 455–463.
irradiation, J. Electroanal. Chem. 765 (2016) 188–196. [183] N. Mojumder, S. Sarker, S.A. Abbas, Z. Tian, V. Subramanian, Photoassisted
[157] W. Wang, F. Li, D. Zhang, D.Y.C. Leung, G. Li, Photoelectrocatalytic hydrogen enhancement of the electrocatalytic oxidation of formic acid on platinized
generation and simultaneous degradation of organic pollutant via CdSe/TiO2 TiO2 nanotubes, ACS Appl. Mater. Interface 6 (2014) 5585–5594.
nanotube arrays, Appl. Surf. Sci. 362 (2016) 490–497. [184] S.S. Shinde, C.H. Bhosale, K.Y. Rajpure, Photodegradation of organic
[158] J. Liu, L. Ruan, S.B. Adeloju, Y. Wu, BiOI/TiO2 nanotube arrays, a unique pollutants using N-titanium oxide catalyst, J. Photochem. Photobiol. B: Biol.
flake-tube structured p-n junction with remarkable visible-light 141 (2014) 186–191.
photoelectrocatalytic performance and stability, Dalton Trans. 43 (2014) [185] J. Pan, X. Li, Q. Zhao, T. Li, M. Tade, S. Liu, Construction of Mn0.5 Zn0.5 Fe2 O4
1706–1715. modified TiO2 nanotube array nanocomposite electrodes and their
[159] X. Zhao, H. Liu, J. Qu, Photoelectrocatalytic degradation of organic photoelectrocatalytic performance in the degradation of 2,4-DCP, J. Mater.
contaminant at hybrid BDD-ZnWO4 electrode, Catal. Commun. 12 (2010) Chem. C 3 (2015) 6025–6034.
76–79. [186] S.Y. Yang, W. Choi, H. Park, TiO2 nanotube array photoelectrocatalyst and
Ni-Sb-SnO2 electrocatalyst bifacial electrodes: a new type of bifunctional
S. Garcia-Segura, E. Brillas / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 31 (2017) 1–35 35

hybrid platform for water treatment, ACS Appl. Mater. Interface 7 (2015) [197] Y.-F. Su, G.-B. Wang, D.T.F. Kuo, M.-L. Chang, Y.-H. Shih,
1907–1914. Photoelectrocatalytic degradation of the antibiotic sulfamethoxazole using
[187] J. Su, L. Zhu, P. Geng, G. Chen, Self-assembly graphitic carbon nitride TiO2 /Ti photoanode, Appl. Catal. B: Environ. 186 (2016) 184–192.
quantum dots anchored on TiO2 nanotube arrays: an efficient [198] L. Liu, R. Li, Y. Liu, J. Zhang, Simultaneous degradation of ofloxacin and
heterojunction for pollutants degradation under solar light, J. Hazard. Mater. recovery of Cu(II) by photoelectrocatalysis with highly ordered TiO2
316 (2016) 159–168. nanotubes, J. Hazard. Mater. 308 (2016) 264–275.
[188] X. Zhao, T. Xu, W. Yao, C. Zhang, Y. Zhu, Photoelectrocatalytic degradation of [199] Z. Hua, Z. Dai, X. Bai, Z. Ye, P. Wang, H. Gu, X. Huang, Copper nanoparticles
4-chlorophenol at Bi2 WO6 nanoflake film electrode under visible light sensitized TiO2 nanotube arrays electrode with enhanced
irradiation, Appl. Catal. B: Environ. 72 (2007) 92–97. photoelectrocatalytic activity for diclofenac degradation, Chem. Eng. J. 283
[189] S. Nissen, B.D. Alexander, I. Dawood, M. Tillotson, R.P.K. Wells, D.E. Macphee, (2016) 514–523.
K. Killham, Remediation of a chlorinated aromatic hydrocarbon in water by [200] X. Cheng, Q. Cheng, X. Deng, P. Wang, H. Liu, A facile and novel strategy to
photoelectrocatalysis, Environ. Pollut. 157 (2008) 72–76. synthesize reduced TiO2 nanotubes photoelectrode for photoelectrocatalytic
[190] M. Fan, C. Yang, W. Pu, J. Zhang, Liquid phase deposition of ZnO film for degradation of diclofenac, Chemosphere 144 (2016) 888–894.
photoelectrocatalytic degradation of p-nitrophenol, Mater. Sci. Semicond. [201] H. Selcuk, J.J. Sene, H.Z. Sarikaya, M. Bekbolet, M.A. Anderson, An innovative
Process. 17 (2014) 104–109. photocatalytic technology in the treatment of river water containing humic
[191] H.C.A. Valdez, G.G. Jiménez, S.G. Granados, C. Ponce de León, Degradation of substances, Water Sci. Technol. 49 (2004) 153–158.
paracetamol by advanced oxidation processes using modified reticulated [202] S.S. Shinde, C.H. Bhosale, K.Y. Rajpure, Hydroxyl radical’s role in the
vitreous carbon electrodes with TiO2 and CuO/TiO2 /Al2 O3 , Chemosphere 89 remediation of wastewater, J. Photochem. Photobio. B: Biol. 116 (2012)
(2012) 1195–1201. 66–74.
[192] C. Liu, D. Fu, H. Li, Behaviour of multi-component mixtures of tetracyclines [203] T. Fang, L. Liao, X. Xu, J. Peng, Y. Jing, Removal of COD and colour in real
when degraded by photoelectrocatalytic and electrocatalytic technologies, pharmaceutical wastewater by photoelectrocatalytic oxidation method,
Environ. Technol. 33 (2012) 791–799. Environ. Technol. 34 (2013) 779–786.
[193] R. Daghrir, P. Drogui, I. Ka, M.A. El Khakani, Photoelectrocatalytic [204] K. Li, H. Zhang, T. Tang, Y. Wang, Y. Xu, D. Ying, J. Jia, Optimization and
degradation of chlortetracycline using Ti/TiO2 nanostructured electrodes application of TiO2 /Ti-Pt photo fuel cell (PFC) to effectively generate
deposited by means of a Pulsed Laser Deposition process, J. Hazard. Mater. electricity and degrade organic pollutants simultaneously, Water Res. 62
199–200 (2012) 15–24. (2014) 1–10.
[194] X. Nie, J. Chen, G. Li, H. Shi, H. Zhao, P.-K. Wong, T. An, Synthesis and [205] X. Zhou, Y. Zheng, J. Zhou, S. Zhou, Degradation kinetics of
characterization of TiO2 nanotube photoanode and its application in photoelectrocatalysis on landfill leachate using codoped TiO2 /Ti
photoelectrocatalytic degradation of model environmental pharmaceuticals, photoelectrodes, J. Nanomater. (2015), ID 810579, 11p.
J. Chem. Technol. Biotechnol. 88 (2013) 1488–1497. [206] J.C. Cardoso, G.C. Bessegato, M.V.B. Zanoni, Efficiency comparison of
[195] X. Cheng, H. Liu, Q. Chen, J. Li, P. Wang, Enhanced photoelectrocatalytic ozonation photolysis, photocatalysis and photoelectrocatalysis methods in
performance for degradation of diclofenac and mechanism with TiO2 real textile wastewater decolorization, Water Res. 98 (2016) 39–46.
nano-particles decorated TiO2 nano-tubes arrays photoelectrode,
Electrochim. Acta 108 (2013) 203–210.
[196] R. Li, S.E. Williams, Q. Li, J. Zhang, C. Yang, A. Zhou, Photoelectrocatalytic
degradation of ofloxacin using highly ordered TiO2 nanotube arrays,
Electrocatalysis 5 (2014) 379–386.

You might also like