PTH Nanocomposite

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

Accepted Manuscript

Polythiophene nanocomposites for photodegradation applications: Past, present


and future

Mohd Omaish Ansari, Mohammad Mansoob Khan, Sajid Ali Ansari, Moo Hwan
Cho

PII: S1319-6103(15)00082-4
DOI: http://dx.doi.org/10.1016/j.jscs.2015.06.004
Reference: JSCS 747

To appear in: Journal of Saudi Chemical Society

Received Date: 28 February 2015


Revised Date: 9 June 2015
Accepted Date: 15 June 2015

Please cite this article as: M.O. Ansari, M.M. Khan, S.A. Ansari, M.H. Cho, Polythiophene nanocomposites for
photodegradation applications: Past, present and future, Journal of Saudi Chemical Society (2015), doi: http://
dx.doi.org/10.1016/j.jscs.2015.06.004

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Polythiophene nanocomposites for photodegradation applications:

Past, present and future

Mohd Omaish Ansari,a Mohammad Mansoob Khan,ab* Sajid Ali Ansari,a and Moo Hwan Choa*
a
School of Chemical Engineering, Yeungnam University, Gyeongsan-si, Gyeongbuk 712-749,
South Korea, Phone: +82-53-810-2517; Fax: +82-53- 810-4631.
b
Chemical Sciences, Faculty of Science, Universiti Brunei Darussalam, Jalan Tungku
Link, BE1410, Brunei Darussalam.


ϭ

ACCEPTED MANUSCRIPT

Polythiophene nanocomposites for photodegradation applications:

Past, present and future

Mohd Omaish Ansari,a Mohammad Mansoob Khan,ab* Sajid Ali Ansari,a and Moo Hwan Choa*
a
School of Chemical Engineering, Yeungnam University, Gyeongsan-si, Gyeongbuk 712-749,
South Korea, Phone: +82-53-810-2517; Fax: +82-53- 810-4631.
b
Chemical Sciences, Faculty of Science, Universiti Brunei Darussalam, Jalan Tungku
Link, BE1410, Brunei Darussalam.

Abstract

Polythiophene (PTh) has been the subject of considerable interest because of its good

environmental stability, unique redox electrical behavior, stability in doped and neutral states,

ease of synthesis, and wide range of applications in many fields. Apart from its applications in

the electrical or electronic field, PTh has shown promising applications in photocatalytic

degradation. The fabrication of a catalyst, metal oxides with PTh, extends the absorption range

of the modified composite system, thereby enhancing the photocatalytic activity under UV or

visible light irradiation. Substituted PTh, such as alkyl substitution, modifies the electronic

properties of the polymer, thereby enlarging the potential for industrial applications. PTh or

substituted PTh when combined with metal, metal oxide or a combination of both, can exhibit

tailorable photocatalytic properties. This review focuses on the chemistry of the band gap

engineering of PTh or PTh based systems and the mechanism of photocatalytic degradation. The

major developments in the field of UV and visible light-assisted photocatalysis are discussed in

terms of the parameters that affect the photocatalytic efficiency. On the other hand, some

Ϯ

ACCEPTED MANUSCRIPT

challenges will still need to be investigated experimentally, which will be mentioned as the scope

for future studies. For simplicity, the review has been classified under a major subheading

depending on the type of composite system used for photocatalysis.

Keywords: Photodegradation; Visible light; UV irradiation; Polythiophene nanocomposites

ϯ

ACCEPTED MANUSCRIPT

Contents

1. Introduction

2. Polythiophene

2.1 Chemical states and band gap of polythiophene

2.2 Chemistry of photocatalytic activity of polythiophene /metal-metal oxide composite and

UV or visible light activity

2.3 Polythiophene composite catalyst used in the field of photodegradation

2.4 Substituted polythiophene composite photocatalyst used in the field of

photodegradation

2.5 Multicomponent polythiophene composite catalyst used in the field of

photodegradation

2.6 Polythiophene based composite for the photocatalytic disinfection of microorganisms

3. Summary, potential and scope for future work

ϰ

ACCEPTED MANUSCRIPT

Abbreviations used

Valence band VB
Conduction band CB
Highest occupied molecular orbital HOMO
Lowest unoccupied molecular orbital LUMO
Ultraviolent UV
Methyl orange MO
Rhodamine B RhB
Rhodamine 6G RhG
Diffuse reflectance spectroscopy DRS
pH at point of zero charge pHPZC
Polyaniline Pani
Polypyrrole PPy
Polythiophene PTh
Poly(3-thiopheneacetic acid) PTh-Ac
Poly(3-octylthiophene-2,5-diyl) POTh
Poly(3-hexylthiophene) P3Th
Poly(3,4-propylenedioxythiophene) PPTh
Poly(fluorene-co-thiophene) PFTh
Poly(3-hexylthiophene-2,5-diyl) PHTh
Photoluminescence PL
Brunauer-Emmett-Teller BET
Layered double hydroxides shells LDHs

ϱ

ACCEPTED MANUSCRIPT

1. Introduction

Owing to the rapid industrialization in recent decades, a huge amount of toxic effluent is

being discharged into various water bodies on a daily basis. These ongoing processes pose a

serious problem for the availability of safe water for drinking, household uses, agriculture,

farming, etc. Therefore, there appears to be an intimate shortage of clean water supply, which

highlights the urgent need for the purification of water, making waste water treatment an

important issue of concern (Shih et al., 2015).

A wide range of methods and technologies have been used to remove organic or inorganic

pollutants from water and waste-water to reduce their impact on the environment. These

methodologies involve adsorption on organic or inorganic materials (Kumar et al., 2013),

photocatalytic degradation, oxidation processes, microbiological, or enzymatic decomposition

(Pricelius et al., 2007). Of these, semiconductor photocatalysis has been widely applied as a

‘‘green’’ technology to the purification of air and the elimination of organic contamination of

water, and has become one of the most important applied facets of heterogeneous catalysis (Cao

et al., 2014).

Semiconducting metal oxides nanoparticles have long been explored as a photocatalytic

material for the degradation of pollutants. TiO2 has attracted considerable attention since

Fujoshima and Honda (Fujishima et al., 1972) discovered its photocatalytic properties in their

water splitting experiment. Similarly many other, metal oxide nanoparticles, such as ZnO

(Rashad et al., 2014; Ansari et al., 2013a), TiO2 (Kalathil et al., 2013; Khan et al., 2013; Khan et

al., 2014), etc. (Ansari et al., 2014b; Ansari et al., 2014c), have been used to degrade non-

biodegradable pollutants via photocatalytic routes.

ϲ

ACCEPTED MANUSCRIPT

On the other hand, the wide band gap of metal oxides (Eg ~3.2 eV or greater) limits their

ability to absorb visible light (λ > 380 nm), which limits their widespread use. Therefore, intense

research has been conducted to lower their band gap, such as with metals or nonmetal doping

(Wang et al., 2014b; Ahmmad et al., 2010), composite synthesis with polymers (Mukherjee et al.,

2014) etc. Among these processes, fabrication with conducting polymers, such as PTh, have

been shown to effectively lower the band gap by allowing greater adsorption in the visible region.

Accordingly, there has been considerable research on increasing the degradation rate of

pollutants by combining inorganic materials with PTh to achieve synergetic and complementary

behaviors between PTh and other organic or inorganic materials.

As a result, a large number of papers on PTh nanocomposite based photocatalysts have been

published within a short span of time. The present review summarizes the systematic progress in

the development, mechanistic view and future applications of PTh-based photocatalytic systems.

For convenience, the discussion is divided into a several categories depending on the type of the

composite material and its chemistry.

2. Polythiophene

Since the discovery of iodine-doped polyacetylene in 1977 (Shirakawa et al., 1977), research

in the field of conducting polymers has made considerable advances. On the other hand, only

few conducting polymers have been found to be stable enough under normal processing

conditions to be incorporated in practical applications. Among them, the leading candidates are

Pani, PPy and PTh. Of these, Pani and PPy have been studied the most for many practical

applications, such as sensors (Ansari et al., 2011; Patois et al., 2012), photocatalytic activity

(Ansari et al., 2014a), environment remediation (Kumar et al., 2013) etc. Despite being the least

ϳ

ACCEPTED MANUSCRIPT

studied, PTh has shown many promising applications comparable to both Pani and PPy, such as

electrical/electrochemical applications, sensors and dye degradation (Gonçalves et al., 2012).

Recently, PTh, just like other conjugated polymers, has been applied in the photocatalytic area to

sensitize metal oxides and develop high performance PTh/metal oxides photocatalytic materials

(Qiu et al., 2012).

In addition to pure PTh, functionalized PTh has also attracted considerable interest owing to

their interesting electrical, electrochemical and optical properties. The introduction of flexible

pendant chains onto the backbone improves the solubility and processability allowing more

complete characterization of the materials. Poly(3-alkylthiophene)s with large alkyl groups, such

as butyl, can be readily melt-or solution-processed into films, which, after oxidation, can exhibit

exceptionally high electrical conductivity (McCullough et al., 1998). Moreover, such

introduction modifies the electronic properties of the polymer increasing the possibilities for

industrial applications. Fig. 1 shows the basic structures of PTh and some of its derivatives.

Poly(3-alkylthiophene)

* S n
*

R
Poly(3-alkylthiophene)

ϴ

ACCEPTED MANUSCRIPT

O O

Poly(3,4-ethylene dixoxythiophene)

OH
Poly(hydroxymethyl-3,4-

ethylenedioxythiophene)

O O

Poly-(fluorene-co-thiophene)

Figure 1. Structures of PTh and its derivatives.

2.1 Chemical states and band gap of polythiophene

PTh exhibits a large range and ‘‘well-behaved’’ chemical properties, making it suitable for

selective and efficient modification at the molecular level that can be used to tailor its physical,

optical and electronic properties. This backbone of PTh can be modified by the incorporation of

different monomer building blocks, such as fused ring systems, and substitution by alkyl chains

ϵ

ACCEPTED MANUSCRIPT

into PTh. This alters the chemical behavior, which in turn affects the band gap, an important

parameter in the field of photocatalysis (Krüger et al., 2011). The band gap of PTh is in the range

of 2.0 eV and varies according to the type of doping. This band gap is sufficiently small

compared to metal oxides, such as TiO2, ZnO etc. Owing to this sufficiently small band gap, it is

possible to excite an electron from the VB to the CB using both UV and visible light. Fig 2

shows the comparative visible and UV light transition possibilities of PTh along with those of

metal oxides.

(a) (b)
CB ē CB ē π∗
LUMO

UV excitation Visible no excitation Eg ∼ 2.5 eV excitation


Eg ∼ 3.2 eV Eg ∼ 3.2 eV UV and Visible
light light
VB h+ VB h+ light HOMO π
Pure metal oxide Pure metal oxide
nanostructure PTh
nanostructure

Figure 2. Different transition possibilities in UV and visible light for (a) metal oxides and (b)

PTh.

The substitution of PTh with either electron withdrawing or electron releasing groups can be

an alternative way for band gap engineering. This tunes the HOMO and LUMO of the

conjugated polymer system, thereby altering the band gap. The introduction of electron

withdrawing groups, such as nitro or cyano at the position of the PTh system, has been reported

to increase the oxidation potential, and affect the band gap. Lambert et al. (Lambert et al., 1991)

reported that poly(cyclopentabithiophene)s bearing electron withdrawing keto or dicyano groups

at the bridging carbon has a band gap of 1.20 eV and 0.80 eV, respectively.

This low band gap has been explained by the increased quinoid character of PTh or its

derivatives (Toussaint et al., 1993). Another assumption is that the systems can also be viewed as
ϭϬ

ACCEPTED MANUSCRIPT

alternant donor acceptor systems, where the thiophene ring acts as a donor. Therefore, the

introduction of electron withdrawing groups may lower the band gap of the PTh system, but have

a serious impact on the solubility, which ultimately decreases their processability and limits their

practical applications (Roncali et al., 2007).

The electron donating group increases the HOMO level of the conjugated system which

reduces the band gap. Alkyl groups, as in the case of butyl or pentyl substituted PTh, decrease

the oxidation potential of the thiophene ring due to the inductive effect (Roncali et al., 1987). On

the other hand, linear alkyl chains of sufficient length (typically 6-9 carbons) contribute

indirectly to reducing the energy gap by enhancing the long-range order in the polymer through

lipophilic interactions between the alkyl chains, which leads to a decrease in the band gap

(Roncali et al., 1987). Compared to electron withdrawing groups, electron donating groups, such

as the long alkyl chains, opens up the structure of the alkyl group substituted PTh or their

derivatives, which increases their solubility, processability and finds a variety of practical

applications.

2.2 Chemistry of the photocatalytic activity of polythiophene /metal-metal oxide composite and

UV or visible light activity

PTh like other conducting polymers under visible light irradiation, absorbs light to induce a

π → π* transition, transporting the excited-state electrons to the π orbital, which is also referred

to as the VB to CB or HOMO to LUMO transition (Zhai et al., 2004). UV and visible irradiation

is sufficient in energy for this transition because of the low band gap of PTh. Therefore, due to

the ease of the electron transition, PTh can act easily as a sensitizer to metal oxides where the

transition of the electron to the CB is not possible under visible light because of its wide band

ϭϭ

ACCEPTED MANUSCRIPT

gap. In the case of metal oxide composites with PTh, the transition of an electron to the CB of

TiO2 occurs with ease due to the polymer-mediated process. Electrons from the CB of PTh are

transferred directly to the CB of metal oxides making the composite active in visible light (Fig.

3). In the case of UV irradiation, the transfer of electrons occurs at a much higher rate due to the

activity of the metal oxide, which increases the catalytic activity considerably. The excited-state

electrons readily injected into the CB of metal oxides are subsequently transferred to the surface

and react with adsorbed water and oxygen to yield hydroxyl and superoxide radicals. These

oxidative radicals are responsible for the photocatalytic activity of the composite of PTh with

metal oxides (Sivakumar et al., 2014).

CB ē
LUMO π∗
Visible Visible
Eg ∼ 2.5 eV excitation no excitation
light Eg ∼ 3.2 eV
light
HOMO π VB h+
Pure metal oxide
PTh
nanostructure

O2
ē
LUMO π∗ ē
CB O2

Visible
Eg ∼ 2.5 eV Visible
light
HOMO light
π VB
PTh Pure metal oxide
nanostructure

Figure 3. PTh mediated electron transfer to the CB of metal oxides under visible light irradiation.

2.3 Polythiophene composite catalyst used in the field of photodegradation

ϭϮ

ACCEPTED MANUSCRIPT

The major drawback of any metal oxide-mediated photocatalysis process is the high rate of

electron and hole recombination, which decreases the overall photocatalysis efficiency. Dye

sensitized photocatalytic materials show high efficiency in the degradation of pollutants, where

the dye acts as a sensitizer and facilitates electron transfer to the CB of TiO2. The electron is

captured to produce active radicals, such as .O2-, .OOH and .OH, which are responsible for the

photodegradation process (Wu et al., 1998; Zhang et al., 2008). On the other hand, the loose

attachment of dyes with the metal oxide is a major drawback, which results in poor stability for

long term applications (Wang et al., 2003). Conducting polymers, such as PTh in the place of

dyes, is an efficient method of resolving these problems because the polymer can be coated on

the surface of metal oxides, and the resulting formed composite would exhibit strong interactions

as well as show high absorption in the UV and visible region of the solar spectrum.

In situ polymerization techniques have mostly been used for the fabrication of metal oxides.

Zhu et al. (Zhu et al., 2008) exploited the properties of the high electron mobility of PTh to

fabricate a PTh/TiO2 nanocomposite for the photocatalytic degradation of MO dye under UV

light irradiation. A self-designed UV-irradiation instrument with 10 W germicidal lamps with a

maximum wavelength of 253.7 nm was used for the catalysis process. The change in the MO

concentration was measured by a UV-vis spectrophotometer and was related to the degradation

of MO. Morphological analysis showed that during in-situ polymerization, PTh is coated on the

TiO2 nanoparticles and the overall composite absorbs in the wavelength range of 200-800 nm

compared to TiO2, which absorbs only in the 400 nm region. This suggests that the PTh/TiO2

nanocomposite can be excited by solar light. In addition to the photodegradation phenomenon,

the PTh/TiO2 nanocomposite also allows surface adsorption of the dye, which changes with the

PTh to TiO2 weight ratio. With the increase in PTh ratio, the adsorption capacity of the PTh/TiO2

ϭϯ

ACCEPTED MANUSCRIPT

nanocomposites increases first and decreases afterwards, which is related to the structural

features of the nanocomposite system. When tested under UV light illumination, the pure PTh

used as a catalyst did not show any appreciable degradation of MO (< 5.0%). Pure TiO2 induced

48.5% MO degradation, whereas PTh/TiO2 showed 56.6% degradation. The rate of degradation

initially starts at a high pace after which it becomes stable due to the adsorption of dye on the

active sites of the catalyst. PTh/TiO2 with a higher PTh content exhibited limited catalytic

activity, which is because PTh adsorbed on the surface of TiO2 occupied most of the surface

active sites of TiO2, thereby preventing the MO molecule from reaching the active sites of TiO2

for degradation.

The surface property is an important factor for the photocatalytic properties and is closely

related to the PTh to TiO2 weight ratio. In another development reported previously, it was

shown that with increasing PTh content from 0 to 2%, the adsorption capacity of the PTh/TiO2

composites towards MO increases (Li et al., 2008). Zeta potential studies showed that in the

dispersion, the surface of the TiO2 is negatively charged while the surface of the PTh//TiO2

composites is positively charged. Therefore, due to coulombic repulsion, MO is barely adsorbed

on the TiO2 surface. On the other hand, the positively charged surface of the PTh/TiO2

composites can adsorb MO strongly, which would contribute to the high photodegradation

efficiency because it facilitates efficient interfacial charge transfer. The adsorbed MO molecules

are also helpful in preventing the recombination of photoelectrons and holes (Xu et al., 2010).

Therefore, the strong adsorption properties of the PTh/TiO2 composites could increase the

photocatalytic degradation rate of MO (Zhu et al., 2010a). The PTh/TiO2 nanocomposite, when

studied for its visible light activity, also showed substantial visible light photocatalytic activity.

TiO2 is excited in by UV light only; hence PTh acts as a sensitizer. PTh because of its strong

ϭϰ

ACCEPTED MANUSCRIPT

absorbance in the visible light region upon excitation transfers the electron to the CB of TiO2.

The CB electron then reacts with O2 adsorbed on the TiO2 surface to produce oxidative radicals

responsible for the photodegradation process.

In addition to the conventional in situ polymerization techniques used for the preparation of

PTh based photocatalysts, other techniques, such as microemulsion route and the electrochemical

or photoinduced polymerization of PTh, also produced interesting results due to the different

morphological features that gives them different properties (Wang et al., 2013). The

microemulsion phase acts as an inert medium to prevent the coalescence of nanoparticles.

Degradation studies using PTh nanoparticles for commonly used textile dyes, such as Orange II

and MO, under UV irradiation indicated higher catalytic activity due to the low band gap of the

PTh nanoparticles. Under UV irradiation, PTh nanoparticles exhibited higher degradation rates

than normal PTh because of the larger specific surface areas. The photocatalytic degradation

reaction simply followed pseudo first-order reaction kinetics. The rate constants of the normal

PTh nanoparticles were relatively low compared to TiO2 and the degradation rate of the PTh

nanoparticles was much higher than TiO2. One explanation is that the energy of the photo-

induced radicals coming from the narrower bang gap (~2.0 eV) of PTh nanoparticles is much

lower than that from TiO2 (~3.2 eV) (Cheng et al., 2012).

The electropolymerization of thiophene on TiO2 films leads to a slow and intact coating of

PTh on TiO2. Liang et al. (Liang et al., 2009) used the electropolymerization technique for the

surface coating of PTh on TiO2. They reported that the low potential avoids degradation of the

polymer as well as any side reactions between the electrolytes and electrodes. The thickness of

the PTh layer on TiO2 can be controlled easily by the concentrations of thiophene monomers in

the electrolyte solution. In the absorption spectra, the PTh/TiO2 composites revealed a broad

ϭϱ

ACCEPTED MANUSCRIPT

absorption band in the range, 400-600 nm, with a main shoulder at 540 nm. This confirmed the

interaction of PTh with TiO2 as the composite processed all the absorption regions of TiO2 and

PTh. Owing to the high absorption in the UV and visible light region, the electropolymerized

PTh/TiO2 composites showed high photocatalytic degradation of 2,3-dichlorophenols when

tested under UV and visible light. The enhanced effect was explained by the high rate of charge

transfer from the sulfur site of PTh to TiO2, as evident from XPS analysis.

A similar interactions of the sulfur site with TiO2 was also achieved in other controlled

polymerization conditions. Under normal conditions, PTh is simply deposited on the surface of

TiO2; hence, the interface between the polymer and TiO2 is not stabilized for the formation of a

heterojunction, which favors the rate of charge injection when stimulated by light excitation.

Photoinduced polymerization is another technique of composite preparation by utilizing

photoexcited TiO2 nanoparticles and achieving the free radical polymerization of alkenes. This

technique leads to good adhesion between the polymeric materials and inorganic materials

because the polymerization reaction is initiated locally by charge transfer across the inorganic

semiconductor-monomer interface (Strandwitz et al., 2010; Weng et al., 2008). The PTh/TiO2

nanocomposite prepared using the photoinduced polymerization technique showed a strong

interaction between the sulfur sites of the PTh backbone and the TiO2 particles, which can

explain the highly efficient photocatalytic degradation of RhB under UV irradiation or visible

light irradiation (Xu et al., 2011).

Apart from TiO2, other materials, such as ZnO and metal-organic frameworks, (MIL-101(Cr)

and Bi25FeO40), in combination with PTh have also shown promising catalytic properties. ZnO

absorbs a larger fraction of the UV spectrum than TiO2 but has limited applications because of

the occurrence of photocorrosion and the susceptibility to facile dissolution at extreme pH (Fu et

ϭϲ

ACCEPTED MANUSCRIPT

al., 2008). Khatamian et al. (Khatamian et al., 2014) reported that the surface modification of

ZnO with PTh can improve the absorption characteristics of ZnO to produce a highly visible

light active photocatalyst. The optical properties of the prepared PTh/ZnO by DRS showed a

broad peak in the visible region. The pHPZC changes with the addition of PTh to ZnO due to a

change in the surface property of the composite. This is also evident from morphological studies

because ZnO nanoparticles are surrounded by PTh, which reduces the agglomeration of ZnO

nanoparticles in the PTh/ZnO system. The BET results showed that the surface area of PTh/ZnO

is higher than that of ZnO and the obtained surface areas of PTh/ZnO and ZnO were 199.74 and

9.9 m2g-1, respectively. A study of pHPZC suggests that the surface of the composite is positively

charged, which is in contrast to ZnO, and MO being negatively charged can be adsorbed easily

on the surface, which is beneficial for the photodegradation properties. The photocatalytic

mechanism has been explained in a similar manner to the case of PTh/TiO2 as discussed earlier,

i.e. the transfer of an electron to the CB of TiO2. On the other hand, the authors concluded that a

specific threshold of PTh is necessary to prevent recombination, and PTh in very small amounts

resulted in poor efficiency. This might be due to the inability of a small amount of PTh to

transfer sufficient electrons to the CB of ZnO, which is further responsible for the degradation

process. Similar characteristics, as in the case of PTh/ZnO have been observed for the

PTh/Bi25FeO40/MIL-101(Cr) composite (Lv et al., 2014). The visible light absorption increased

due to the high absorption of MIL-101(Cr). Therefore, the catalysis of RhB was achieved

efficiently under visible light irradiation. The photocatalytic mechanism is believed to occur via

a polymer mediated process, i.e. transfer and separation of photogenerated charge carriers

through the PTh-connected interface between MIL-101 and Bi25FeO40.

ϭϳ

ACCEPTED MANUSCRIPT

2.4 Substituted polythiophene composite photocatalyst used in the field of photodegradation

Derivatives of PTh, such as alkyl substitution, improves the processability of the polymer

and may extend their applications to various fields. Among the different substituted PTh, the

alkyl substituted P3Th has been applied widely in the field of photocatalytic degradation. The

alkyl group acts as an electron donor, which is helpful in increasing the effective conjugation

length of thiophene ring backbone and enhancing the electron mobility. P3Th has strong

absorption in visible region (band gap of 1.9-2.0 eV), higher charge carrier mobility, good

solubility, processability, long-term stability (Motaung et al., 2009). Therefore, it can be

considered a good candidate as a sensitizer of TiO2 and may be a possible replacement for PTh

due to the higher processability of the former.

P3Th can be synthesized using 3-hexylthiophene as a monomer using a simple in situ

oxidative polymerization technique similar as in the case of PTh. The composite of P3Th/TiO2

exhibits broad absorption, both in the UV and visible regions of the solar spectrum. Muktha et al.

(Muktha et al., 2007) from their PL results, reported a quenching phenomenon as the content of

TiO2 in the composites was increased, which was attributed to the presence and distribution of

TiO2 in the P3Th matrix because charge transfer occurs from P3Th to TiO2. They explained that

upon UV excitation, the electron from the CB of TiO2 may be excited to the empty states in

P3Th owing to the lower band gap of P3Th compared to that of TiO2. The process enhances the

oxidative property of TiO2 and stabilizes the VB hole of TiO2. On the other hand, P3Th may also

absorb UV radiation allowing energy transfer to the TiO2 nanoparticles followed by the creation

of a VB hole and CB electron. This electron-mediated mechanism leads to the formation of ͻOH

radicals responsible for the degradation process.

ϭϴ

ACCEPTED MANUSCRIPT

Similar studies have been carried out by various authors to examine the absorption and

visible light activity of P3Th/TiO2 nanocomposites. In a study by D. Wang et al. (Wang et al.,

2009), DRS spectroscopy showed that P3Th/TiO2 possesses a much lower band gap than TiO2

and can be easily excited to produce more electron-hole pairs under visible light irradiation.

They reported that the optimal P3Th to TiO2 ratio gave the highest photocatalytic efficiency. The

catalytic degradation rate of the MO dye increased initially and later decreased with increasing

P3Th/TiO2 molar ratio from 30:1 to 125:1, reaching a maximum of 88.5% in 10 h when the

molar ratio of P3Th/TiO2 was 75:1. The photocatalytic mechanism was found to be similar, via

electron transfer from P3Th to the CB of TiO2. The electron in the CB of TiO2 reacts with

oxygen to yield a superoxide radical, ͻO2-. This leads to the formation of a positively charged

hole (h+) due to electron migration from the VB of TiO2 to the π orbital of P3Th, which then

reacts with OH- or H2O to generate .OH. The super-oxide radical ion (O2-) and hydroxyl radical

(ͻOH) are responsible for the degradation of the organic compounds.

In another development, Zhu et al. (Zhu et al., 2010b), who used P3Th/TiO2 for the

degradation of MO dye, reported that the catalytic reactions occurred via different reactive

pathways under UV and visible light irradiation. The azo bond of the MO molecule was

dominantly cleaved under UV light irradiation, while under visible irradiation, apart from the

cleavage of the azo bond, intermediates with chromophoric groups formed. Further studies on

the degree of P3Th oxidation showed that the excessive oxidation of P3Th decreases the

photocatalytic activity of the P3Th/TiO2 composites. From the degradation results of MO using

P3Th/TiO2 with different degrees of P3Th oxidation, it was concluded that as the side chains of

P3Th became oxidized, they had a stronger interaction with TiO2, leading to an increase in the

polarity of P3Th (Xu et al., 2012). This reduces the separation ability of P3Th to oxygen, and the

ϭϵ

ACCEPTED MANUSCRIPT

photoinduced electrons can be transferred easily from P3Th to TiO2 particles, making the

P3Th/TiO2 composite photocatalytic active. Under extensive oxidation, however, the backbone

of the polymer was destroyed, leading to a decrease in the degree of π-π conjugation of the

polymer. This decreases the efficiency of the photo-excitation of electrons from the valence bond

to the antibonding polaron state under visible light irradiation. Therefore, fewer electrons can be

excited by visible light irradiation and be transferred to the conducting band of TiO2 to generate

oxidative species. Thus, the photocatalytic activity of the P3Th/TiO2 composites decreases when

P3Th is oxidized excessively.

Another derivative of PTh, PFTh, in combination with TiO2 also showed enhanced catalytic

degradation properties (Song et al., 2007). DRS studies revealed a broad absorption for

PFTh/TiO2 in the region from 360 to 500 nm compared to pure TiO2, where no absorption was

observed after 400 nm. The PL results indicated that the electrons on the CB of PFTh were

transferred to the CB of TiO2. Owing to the lower reductive potential of PFTh (-1.35 V) than

TiO2 (-0.65 V), it became thermodynamically possible for electrons to be injected from the CB

of the polymer to the CB of TiO2, which plays an important role in photocatalytic degradation.

Similar results have been reported for the photocatalytic degradation of phenol, RhB and MO by

PFTh/ZnO under visible light irradiation (Qiu et al., 2008). For the PPTh/ZnO composite, it was

shown that the amount of ZnO in the polymer has a great effect on the conjugation length and

degree of oxidation of the composite (Ali et al., 2013). Among 3, 5 and 7% ZnO loadings in the

polymer, FTIR and UV-visible spectral analysis showed that a 5% loading had the longest

conjugation length compared to other weight percentages. This can be explained in that when

ZnO is excited by radiation capable of electron transfer from the VB to the CB, the electron-hole

pairs react with hydroxyl ions to produce reactive radicals, such as .O2- and .OH (Fu et al., 2012).

ϮϬ

ACCEPTED MANUSCRIPT

These strong oxidants in stepwise reactions can enhance the conjugation length of the polymer.

On the other hand, ZnO at higher weight percentages will also lead to the separation of oxidants

to some extent due to other possible reactions; hence, a higher weight percentage may lead to a

decrease in the conjugation length (Abdiryim et al., 2012). The photocatalytic activity is

explained by the synergistic effect of PPTh and ZnO. The 5% ZnO loading gave the highest

photocatalytic efficiency under UV irradiation for the dye. The higher conjugation and oxidation

degree leads to good electron mobility and is believed to be an important factor for enhancing the

photocatalytic activity.

Little work has also been done on the photocatalytic activity of carbon-based materials with

a PTh composite. GO has a functional oxygen group capable of interacting with PTh or its

derivative via a π-π interaction between the nonpolar regions of GO and the molecular layers of

PTh. The composite PHTh/GO sheets, when studied for their photophysical properties, such as

UV-visible absorption, photothermal desorption spectroscopy, and steady-state PL studies,

showed enhanced optical absorption, charge transfer and photocatalytic properties (Wang et al.,

2012). The formation of a charge-transfer complex between GO and PHTh was proposed, which

creates additional sub-band gap states that result in absorption at lower energies than the band

gap. PHTh/GO hybrids when studied for the oxidative Mannich reaction under ambient solar

light showed promising results. Therefore, owing to its ability to be excited by sunlight,

PHTh/GO hybrids have potential applications as an efficient photocatalyst in organic synthesis.

2.5 Multicomponent polythiophene composite catalyst used in the field of photodegradation

Multicomponent systems consisting of PTh/metal oxides and metal particles have also

attracted attention to further extend the absorption limit of the composite system and enhance the

Ϯϭ

ACCEPTED MANUSCRIPT

catalytic activity (Wang et al., 2014a). Chandra et al. (Chandra et al., 2015) reported that a

PTh/TiO2-Cu composite prepared by a sol-gel process showed a strong interaction between the

interface of PTh and TiO2-Cu. The TiO2-Cu composite showed absorption in the region, 450-

500nm. On the other hand, extension of the absorption regions in the PTh/TiO2-Cu composite

also indicates strong interactions between PTh and TiO2-Cu, which is expected to accelerate

photo-induced charge transfer from the PTh to TiO2-Cu, and subsequent enhancement of the

photocatalytic efficiency of the catalyst. When tested for the degradation of RhB, as the PTh

content in TiO2-Cu was increased from 0.5 to 2.0 wt.%, the rate of RhB degradation increased,

reaching a maximum at polymer loading of 1.0 wt.%, and decreasing with further increases in

polymer loading to 2.0 wt.%. This shows that there is some specific threshold for PTh to exhibit

the optimal catalytic properties in the composite system. The experiment also showed that the

concentration of PTh in TiO2-Cu played a significant role in enhancing the photocatalytic

activity. The change in pH alters the charge on the catalyst, which affects the adsorption and

desorption of the dyes on the catalyst surface. For PTh/TiO2-Cu, the photocatalytic rate also

increased with increasing pH. This is due to the effect of pH on the ionization rate on the surface

according to the following reactions:

TiOH + H+ TiOH2+ (1)

TiOH + OH- TiO- + H2O (2)

The pH study showed that the dispersion of PTh/TiO2-Cu is positively charged under acidic

conditions, which is in contrast to negatively charged TiO2-Cu. Therefore, the adsorption of

positive charged RhB becomes impossible under acidic conditions due to electrostatic repulsion;

however, in alkaline pH, the charge on the catalyst is reversed and the condition becomes

favorable for photocatalytic activity. Another explanation is that the ͻOH radicals are easier to

ϮϮ

ACCEPTED MANUSCRIPT

generate by oxidizing more hydroxide ions on the TiO2 surface; thus, the efficiency of the

photocatalytic process is enhanced in alkaline media.

The photodegradation mechanism proceeds normally via a polymer mediated process;

however the important difference is that upon excitation, the electrons from the VB of TiO2 are

trapped by Cu2+ to form an unstable TiO2-Cu2+ complex. The trapped TiO2-Cu2+ electrons are

moved immediately to the HOMO of PTh, which eliminates their electron-hole recombination

process. Owing to the lower reduction potential of PTh than TiO2-Cu, the electrons are

transferred easily to the CB of TiO2 and a further degradation reaction occurs similarly, as

discussed earlier. Fig. 4 shows a schematic diagram of the photocatalytic degradation mechanism.

O2
ē
LUMO π∗ ē
CB O2

Visible
Eg ∼ 2.5 eV
ē
light Visible
Cu2+ light
HOMO π HO•
VB h+
PTh
Pure metal oxide
nanostructure

Figure 4. Metal mediated electron transfer to PTh and metal oxide.

In another major development, PTh/TiO2-montmorillonite was evaluated for the

sonophotocatalytic degradation of RhG (Boutoumi et al., 2012). When sonication is added to the

photocatalytic system, the transformation of organic products occurs through reactions with .OH

radicals generated from the collapsing bubbles. In addition, the concentration of the organic

material decreases as hydroxylation progresses. The enhanced effects by photocatalysis and

sonocatalysis can be attributed to the following reasons: the increased amounts of the oxidative

Ϯϯ

ACCEPTED MANUSCRIPT

species generated, and an increase in the mass transport of chemical species between the solution

phase and the photocatalyst surface and vice versa (Pirola et al., 2012). PTh/TiO2-

Montmorillonite, when assessed for catalytic reactions, showed a higher sonocatalytic

degradation rate of RhG compared to the photocatalytic process. This was attributed to the

advantageous properties of sonication favoring the ease of generation of charged species as well

as to the effect of PTh, which promoted electron transfer from the excited PTh polymer to the

CB of TiO2. These two effects combined gave high photocatalytic efficiency for the degradation

of RhG.

2.6 Polythiophene based composite for the photocatalytic disinfection of microorganisms

Matsunaga et al. (Matsunaga et al., 2006) reported that the TiO2 photocatalyst may cause the

death of bacterial cells in aqueous media under illumination. Since then, extensive studies have

been carried out on the photocatalytic disinfection of microorganisms using TiO2 as a

photocatalyst under UV light irradiation (Akhavan et al., 2009). The inability of TiO2 to absorb

the visible light, however, has limited its applications. Doping TiO2 with PTh may extend the

absorption region of TiO2 and then enhance the photocatalytic disinfection phenomenon (Čík et

al., 2006). The photocatalytic activity is due to the .OH generated during the photocatalytic

process, which is supposed to be the major lethal species responsible for photocatalytic

disinfection (Shang et al., 2011).

Shang et al. (Shang et al., 2013) in their study of PTh-Ac coated Fe3O4@LDHs magnetic

nanospheres as a photocatalyst for the photocatalytic disinfection of pathogenic bacteria under

solar light irradiation, showed that PTh-Ac plays a key role in the disinfection process. The .OH

radicals that are responsible for photocatalytic disinfection can be generated easily from

Ϯϰ

ACCEPTED MANUSCRIPT

superoxide radicals (•O2-) or singlet oxygen (1O2) on the surface of PTh-Ac. Therefore, it was

assumed that the photocatalytic disinfection mechanism of this study was due to the creation of

OH, which is generated from the surface of Fe3O4@PTh-Ac-LDHs. The generated .OH could

cause significant disorder in the permeability of bacterial cells, DNA damage and decomposition

of the cell walls. Experiments performed on E. coli confirmed that the cell structure was

distorted, and the ruptured cell wall could be observed clearly by TEM analysis. The cell was no

longer intact and the outer membrane of the cell was damaged, leading to leakage of the interior

component. This highlights the substantial disorder in membrane permeability by •OH in the

disinfection process, followed by the free efflux of intracellular constituents, which leads to cell

death. Derivatives of PTh in film form when studied for the disinfection of fungicide also

showed promising applications under visible light. The POTh film, when studied for the visible

light-induced catalytic degradation of iprobenfos fungicide using bubbling gases with varying

oxygen partial pressures, showed a high rate of fungicide degradation (Wen et al., 2000a; Wen et

al., 2000b). The rate of fungicide degradation increased with increasing oxygen partial pressure,

and the film showed remarkable stability, even after 10 h of irradiation (>380 nm), with no

peeling of the films. To find out the active species responsible for the degradation process, the

fluorescence quenching constant (kq) was determined from the slope of the Stern-Volmer plots.

The estimated kq value was 3.6 x 10-3 kPa-1, which is related to the formation of •O2- by electron

transfer from an excited POTh film to molecular oxygen. This formation of •O2- suggests that the

photocatalytic process involves a series of oxidative radicals, such as •O2-, HO2., H2O2, and •OH,

which is similar to the case when TiO2 was used as a photocatalyst (Hoffmann et al., 1995).

Compared to TiO2 particles in aqueous suspensions, the POTh film showed a low quantum yield,

hence, the photocatalytic activity of the polymer film was inferior to that of TiO2 particles (Wen

Ϯϱ

ACCEPTED MANUSCRIPT

et al., 2000a; Wen et al., 2000b). On the other hand, the application of TiO2 particles to the

degradation of wastewater treatments is generally difficult because of their difficult filtration and

settling. The low activity of the polymer film can be attributed mainly to the low redox potential

compared to that of TiO2 as well as to the low surface area. Therefore, the combined effect of

TiO2 with the POTh film can be expected to give enhanced disinfection properties and may also

solve the problems of the settling of TiO2 nanoparticles.

3. Summary, prospective and scope for future works

Many studies of photodegradation under UV and visible light irradiation using composites

based on PTh or its derivatives have been reported. Compared to other conducting polymers,

such as Pani or PPy, PTh has attracted less attention and the research on the photocatalytic

properties of PTh can be considered to be in its infancy at this stage.

Among the metal oxides examined thus far, TiO2 and ZnO have attracted attention in this

area, whereas there are very few reports on other metal oxides. Composites of PTh or its

derivatives can be prepared with different metal oxides to assess their photocatalytic properties

under UV and visible light irradiation. Carbon-based materials, such as graphene, CNT or GO

also have exciting future prospects owing to their conductive nature, and may give high charge

mobility in composites, which is beneficial for the catalytic properties. Owing to the much

broader absorption in the visible region of PTh and carbon based materials, more research will be

needed in the field of visible light photocatalysis, which might provide a green photocatalytic

methodology for future applications.

PTh films can be used for the immobilization of metal oxides, which can solve the problems

of separation and isolation of the catalyst and greatly improve the reproducibility of the

Ϯϲ

ACCEPTED MANUSCRIPT

composite photocatalyst. In addition, there appears to be further scope for the synthesis of PTh

based photocatalysts with different morphologies, such as core shell, wire, network structure etc.

using simple chemical or electrochemical routes. Green synthetic methods, such as catalyst

synthesis via electrochemically active biofilms have attracted considerable interest these days,

and it might be the most common possible methodology for the synthesis of PTh or its

derivative-based photocatalysts in the near future.

Declaration of interest

The authors report no conflicts of interest. The authors alone are responsible for the content and

writing of the paper.

Corresponding Authors : mhcho@ynu.ac.kr; mmansoobkhan@yahoo.com

Acknowledgements

This study was supported by Priority Research Centers Program through the National Research

Foundation of Korea (NRF) funded by the Ministry of Education (2014R1A6A1031189).

References

Abdiryim, T., Ubul, A., Jamal, R., Tian, Y., Awut, T., Nurulla, I., 2012. Solid-state synthesis

and characterization of polyaniline/nano-TiO2 composite. J. Appl. Polym. Sci. 126, 697-

705.

Akhavan, O., Ghaderi, E., 2009. Photocatalytic reduction of graphene oxide nanosheets on

TiO2 thin film for photoinactivation of bacteria in solar light irradiation. J. Phys. Chem. C

113, 20214-20220.

Ϯϳ

ACCEPTED MANUSCRIPT

Ahmmad, B., Kusumoto, Y., Islam, M.S., 2010. One-step and large scale synthesis of non-metal

doped TiO2 submicrospheres and their photocatalytic activity. Adv. Powder Technol. 21,

292-297.

Ali, A., Jamal, R., shao, W., Rahman, A., Osman, Y., Abdiryim, T., 2013. Structure and

properties of solid-state synthesized poly(3,4-propylenedioxythiophene)/nano-ZnO

composite. Prog. Nat. Sci.: Materials international 23, 524-531.

Ansari, M.O., Khan, M.M., Ansari, S.A., Raju, K., Lee, J., Cho, M.H., 2014a. Enhanced thermal

stability under DC electrical conductivity retention and visible light activity of

Ag/TiO2@Polyaniline nanocomposite film. ACS Appl. Mater. Interfaces 6, 8124-8133.

Ansari, M.O., Mohammad, F., 2011. Thermal stability, electrical conductivity and ammonia

sensing studies on p-toluenesulfonic acid doped polyaniline:titanium dioxide

(pTSA/Pani:TiO2) nanocomposites. Sensor Actuat. B-Chem. 157, 122-129.

Ansari, S.A., Khan, M.M., Ansari, M.O., Kalathil, S., Lee, J., Cho, M.H., 2014b. Band gap

engineering of CeO2 nanostructure using an electrochemically active biofilm for visible

light applications. RSC Adv. 4, 16782-16791.

Ansari, S.A., Khan, M.M., Ansari, M.O., Lee, J., Cho, M.H., 2014c. Visible light-driven

photocatalytic and photoelectrochemical studies of Ag–SnO2 nanocomposites

synthesized using an electrochemically active biofilm. RSC Adv. 4, 26013-26021.

Ansari, S.A., Khan, M.M., Kalathil, S., Nisar, A., Lee, J., Cho, M.H., 2013. Oxygen vacancy

induced band gap narrowing of ZnO nanostructures by an electrochemically active

biofilm. Nanoscale 5, 9238-9246.

Ϯϴ

ACCEPTED MANUSCRIPT

Boutoumi, N.K., Boutoumi, H., Khalaf, H., David, B., 2013. Synthesis and characterization of

TiO2-Montmorillonite/Polythiophene-SDS nanocomposites: Application in the

sonophotocatalytic degradation of rhodamine 6G. Appl. Clay Sci. 80-81, 56-62.

Cao, F., Yin, P., Liu, X., Liu, C., Qu, R., 2014. Mercury adsorption from fuel ethanol onto

phosphonated silica gel prepared by heterogenous method. Renew. Energ. 71, 61-68.

Chandra, M.R., Rao, T.S., Pammi, S.V.N., Sreedhar, B., 2015. An enhanced visible light active

rutile titania-copper/polythiophene nanohybrid material for the degradation of rhodamine

B dye. Mat. Sci. Semicon. Proc. 30, 672-681.

Cheng, Y., Wang, J., Bao, Y.C., Ma, Y.L., Wang G.H., 2012. Synthesis of polythiophene

nanoparticles by microemulsion polymerization for photocatalysis

Čík, G., Priesolová, S., Bujdáková, H., Šeršeň, F., Potheöová, T., Krištín, J., 2006. Inactivation

of bacteria G+-S. aureus and G−-E. coli by phototoxic polythiophene incorporated in

ZSM-5 zeolite. Chemosphere 63, 1419-1426.

Fu, D., Han, G., Chang, Y., Dong, J., 2012. The synthesis and properties of ZnO–graphene nano

hybrid for photodegradation of organic pollutant in water. Mater. Chem. Phys. 132, 673-

681.

Fu, H., Xu, T., Zhu, S., Zhu, Y., 2008. Photocorrosion inhibition and enhancement of

photocatalytic activity for ZnO via hybridization with C60. Environ. Sci. Technol. 42,

8064-8069.

Fujishima, A., Honda, K., 1972. Electrochemical photolysis of water at a semiconductor

electrode. Nature 238, 37-38.

Ϯϵ

ACCEPTED MANUSCRIPT

Gonçalves, V.C., Balogh, D.T., 2012. Optical chemical sensors using polythiophene derivatives

as active layer for detection of volatile organic compounds. Sensor Actuat. B-Chem. 162,

307-312

Hoffmann, M.R., Martin, S.T., Choi, W., Bahnemann, D.W., 1995. Environmental applications

of semiconductor photocatalysis. Chem. Rev. 95, 69-96.

Kalathil, S., Khan, M.M., Ansari, S.A., Lee, J., Cho, M.H., 2013. Band gap narrowing of

titanium dioxide (TiO2) nanocrystals by electrochemically active biofilms and their

visible light activity. Nanoscale 5, 6323-6326.

Khan, M.M., Ansari, S.A., Amal, M.I., Lee, J., Cho, M.H., 2013. Highly visible light active

Ag@TiO2 nanocomposites synthesized using an electrochemically active biofilm: a novel

biogenic approach. Nanoscale 5, 4427-4435.

Khan, M.M., Ansari, S.A., Pradhan, D., Ansari, M.O., Han, D.H., Lee, J., Cho, M.H., 2014.

Band gap engineered TiO2 nanoparticles for visible light induced photoelectrochemical

and photocatalytic studies. J. Mater. Chem. A. 2, 637-644.

Khatamian, M., Fazayeli, M., Divband, B., 2014. Preparation, characterization and

photocatalytic properties of polythiophene-sensitized zinc oxide hybrid nanocomposites.

Mat. Sci. Semicon. Proc. 26, 540-547.

Krüger, R.A., Gordon, T.J., Sutherland, T.C., Baumgartner , T., 2011. Band-gap engineering of

polythiophenes. J. Polym. Sci. Pol. Chem. 49, 1201-1209.

Kumar, R., Ansari, M.O., Barakat, M.A., 2013. DBSA doped polyaniline/multi-walled carbon

nanotubes composite for high efficiency removal of Cr(VI) from aqueous solution. Chem.

Eng. J. 228, 748-755.

ϯϬ

ACCEPTED MANUSCRIPT

Lambert, T.M., Ferraris, J.P., 1991. Narrow band gap polymers: polycyclopenta[2,1-b;3,4-

b′]dithiophen-4-one. J. Chem. Soc. Chem. Comm. (11) 752-754.

Liang, H.C., Li, X.Z., 2009. Visible-induced photocatalytic reactivity of polymer–sensitized

titania nanotube films. Appl. Catal-B Environ. 86, 8-17.

Li, Y., Lu, A., Wang, C., Wu, X., 2008. Characterization of natural sphalerite as a novel visible

light-driven photocatalyst. Sol. Energ. Mat. Sol. C 92, 953-959.

Lv, M., Yang, H., Xu, Y., Chen, Q., Liu, X., Wei, F., 2014. Improving the visible light

photocatalytic activities of Bi25FeO40/MIL-101/PTH via polythiophene wrapping. J.

Environ. Chem. Eng. doi:10.1016/j.jece.2014.11.009 (In Press).

Matsunaga, T., Tomoda, R., Nakajima, T., Wake, H., 2006. Photochemical sterilization of

microbial cells by semiconductor powders. FEMS Microbiol. Lett. 29, 211-214.

McCullough, R.D., 1998. The chemistry of conducting polythiophenes. Adv. Mater. 10, 93-116.

Motaung, D.E., Malgas, G.F., Arendse, C.J., Mavundla, S.E., Oliphant, C.J., Knoesen, D., 2009.

Thermal-induced changes on the properties of spin-coated P3HT:C60thin films for solar

cell applications. Sol. Energ. Mat. Sol. C 93, 1674-1680.

Mukherjee, D., Barghi, S., Ray, A.K., 2014. Preparation and characterization of the

TiO2 immobilized polymeric photocatalyst for degradation of aspirin under UV and solar

light. Processes 2, 12-23.

Muktha, B., Mahanta, D., Patil, S., Madras, G., 2007. Synthesis and photocatalytic activity of

poly(3-hexylthiophene)/TiO2composites. J. Solid State Chem. 180, 2986-2989.

Patois,T., Sanchez, J.B., Berger, F., Rauch, J.Y., Fievet, P., Lakard, B., 2012. Ammonia gas

sensors based on polypyrrole films: Influence of electrodeposition parameters. Sensor

Actuat. B Chem. 171-172, 431-439.

ϯϭ

ACCEPTED MANUSCRIPT

Pirola, M.M.C., Selli, E., 2003. Degradation of organic water pollutants through

sonophotocatalysis in the presence of TiO2. Ultrason. Sonochem. 10, 247-254.

Pricelius, S., Held, C., Sollner, S., Deller, S., Murkovic, M., Ullrich, R., Hofrichter, M., Cavaco-

Paulo, A., Macheroux, P., Guebitz, G.M., 2007. Enzymatic reduction and oxidation of

fibre-bound azo-dyes. Enzyme Microb. Tech. 40, 1732-1738.

Qiu, R., Zhang, D., Diao, Z., Huang, X., He, C., Morel, J.L., Xiong, Y., 2012. Visible light

induced photocatalytic reduction of Cr(VI) over polymer-sensitized TiO2 and its

synergism with phenol oxidation. Water Res. 46, 2299-2306.

Qiu, R., Zhang, D., Mo, Y., Song, L., Brewer, E., Huang, X., Xiong, Y., 2008. Photocatalytic

activity of polymer-modified ZnO under visible light irradiation. J. Hazard. Mater. 156,

80-85.

Rashad, M.M, Ismail, A.A., Osama, I., Ibrahim, I.A., Kandil, A.T., 2014. Photocatalytic

decomposition of dyes using ZnO doped SnO2nanoparticles prepared by solvothermal

method. Arabian J. Chem. 7, 71-77.

Roncali, J., Garreau, R., Yassar, A., Marque, P., Garnier, F., Lemaire, M., 1987. Effects of steric

factors on the electrosynthesis and properties of conducting poly( 3-alkyithiophenes). J.

Phys. Chem. 91, 6706-6714.

Roncali, J., 2007. Molecular engineering of the band gap of π-conjugated systems: facing

technological applications. Macromol. Rapid Comm. 28, 1761-1775.

Shang, K., Ai, S., Ma, Q., Tang, T., Yin, H., Han, H., 2011. Effective photocatalytic disinfection

of E. coli and S. aureus using polythiophene/MnO2 nanocomposite photocatalyst under

solar light irradiation. Desalination 278, 173-178.

ϯϮ

ACCEPTED MANUSCRIPT

Shang, K., Sun, B., Sun, J., Li, J., Ai, S., 2013. Poly-(3-thiopheneacetic acid) coated

Fe3O4@LDHs magnetic nanospheres as a photocatalyst for the efficient photocatalytic

disinfection of pathogenic bacteria under solar light irradiation. New J. Chem. 37, 2509-

2514.

Shih, P.K., Chang, W.L., 2015. The effect of water purification by oyster shell contact bed. Ecol.

Eng. 77, 382-390.

Shirakawa, H., Louis, E.J., MacDiarmid, A.G., Chiang, C.K., Heeger, A.J., 1977. Synthesis of

electrically conducting organic polymers: halogen derivatives of polyacetylene, (CH)x. J.

Chem. Soc. Chem. Commun. (16), 578-580.

Sivakumar, K., Kumar, V.S., Shim, J.J., Yuvaraj, H., 2014. Photocatalytic and Antimicrobial

Activities of Poly(aniline-co-o-anisidine)/Zinc Oxide Nanocomposite. Asian J. Chem. 26,

600-606

Song, L., Qiu, R., Mo, Y., Zhang, D., Wei, H., Xiong, Y., 2007. Photodegradation of phenol in a

polymer-modified TiO2 semiconductor particulate system under the irradiation of visible

light. Catal. Commun. 8, 429-433.

Strandwitz, N.C., Nonoguchi, Y., Boettcher, S.W., Stucky, G.D., 2010. In Situ

Photopolymerization of Pyrrole in Mesoporous TiO2. Langmuir 26, 5319-5322.

Toussaint, J.M., Brédas, J.L., 1993. Novel low bandgap polymers: polydicyanomethylene -

cyclopentadithiophene and -dipyrrole. Synt. Met. 23, 103-106

Wang, D., Zhang, J., Luo, Q., Li, X., Duan, Y., An, J., 2009. Characterization and photocatalytic

activity of poly(3-hexylthiophene)-modified TiO2 for degradation of methyl orange under

visible light. J. Hazard Mater. 169, 546-550.

ϯϯ

ACCEPTED MANUSCRIPT

Wang, J., Zhang, D., 2013. One-Dimensional Nanostructured Polyaniline: Syntheses,

Morphology Controlling, Formation Mechanisms, New Features, and Applications. Adv.

Polym. Tech. 32, 323-368.

Wang, P., Zakeeruddin, S.M., Moser, J.E., Nazeeruddin, M.K., Sekiguchi, T., Grätzel, M., 2003.

A stable quasi-solid-state dye-sensitized solar cell with an amphiphilic ruthenium

sensitizer and polymer gel electrolyte. Nature Mater. 2, 402-407.

Wang, S., Nai, C.T., Jiang, X.F., Pan, Y., Tan, C.H., Nesladek, M., Xu, Q.H., Loh, K.P., 2012.

Graphene oxide–polythiophene hybrid with broad-band absorption and photocatalytic

properties. J. Phys. Chem. Lett. 3, 2332-2336.

Wang, Y., Chu, W., Wang, S., Li, Z., Zeng, Y., Yan, S., Sun, Y., 2014a. Simple synthesis and

photoelectrochemical characterizations of polythiophene/Pd/TiO2 composite

microspheres. ACS Appl. Mater. Interface 6, 20197-20204.

Wang, Y., Zhang, R., Li, J., Li, L., Lin, S., 2014b. First-principles study on transition metal-

doped anatase TiO2. Nanoscale Research Letters 9, 46-53.

Wen, C., Hasegawa, K., Kanbara, T., Kagaya, S., Yamamoto, T., 2000a. Photocatalytic

properties of poly(3-octylthiophene-2,5-diyl) film blended with sensitizer for the

degradation of iprobenfos fungicide. J. Photoch. Photobio. A 137, 45-51.

Wen, C., Hasegawa, K., Kanbara, T., Kagaya, S., Yamamoto, T., 2000b. Visible light-induced

catalytic degradation of iprobenfos fungicide by poly(3-octylthiophene-2,5-diyl) film. J.

Photoch. Photobio. A 133, 59-66.

Weng, Z., Ni, X., 2008. Oxidative polymerization of pyrrole photocatalyzed by

TiO2 nanoparticles and interactions in the composites. J. Appl. Polym. Sci. 110, 109-116.

ϯϰ

ACCEPTED MANUSCRIPT

Wu, T., Liu, G., Zhao, J., Hidaka, H., Serpone, N., 1998. Photoassisted degradation of dye

pollutants. V. self-photosensitized oxidative transformation of Rhodamine B under visible

light irradiation in aqueous TiO2 dispersions. J. Phys. Chem. B 102, 5845-5851.

Xu, S., Gu, L., Wu, K., Yang, H., Song, Y., Jiang, L., Dan, Y., 2012. The influence of the

oxidation degree of poly(3-hexylthiophene) on the photocatalytic activity of poly(3-

hexylthiophene)/TiO2 composites. Sol. Energ. Mat. Sol. C 96, 286-291.

Xu, S., Jiang, L., Yang, H., Song, Y., Dan, Y., 2011. Structure and photocatalytic activity of

polythiophene/TiO2 composite particles prepared by photoinduced polymerization.

Chinese J. Catal. 32, 536-545.

Xu, S., Zhu, Y., Jiang, L., Dan, Y., 2010. Visible light induced photocatalytic degradation

of methyl orange by polythiophene/TiO2 composite particles. Water Air Soil Poll.

213, 151-159.

Zhai, L., McCullough, R.D., 2004. Regioregular polythiophene/gold nanoparticle hybrid

materials. J. Mater. Chem. 14, 141-143.

Zhang, X., Sun, D.D., Li, G., Wanga, Y., 2008. Investigation of the roles of active oxygen

species in photodegradation of azo dye AO7 in TiO2 photocatalysis illuminated by

microwave electrodeless lamp. J. Photochem. Photobiol. A 199, 311-315

Zhu, Y., Dan, Y., 2010. Photocatalytic activity of poly(3-hexylthiophene)/titanium dioxide

composites for degrading methyl orange. Sol. Energ. Mat. Sol. C 94, 1658-1664.

Zhu, Y., Xu, S., Jiang, L., Pan, K., Dan, Y., 2008. Synthesis and characterization of

polythiophene/titanium dioxide composites. React. Funct. Polym. 68, 1492-1498.

Zhu, Y., Xu, S., Yi, D., 2010. Photocatalytic degradation of methyl orange using

polythiophene/titanium dioxide. React. Funct. Polym. 70, 282-287.

ϯϱ


You might also like