You are on page 1of 15

Chapter 1

Mechanism of Photocatalysis

1.1 Introduction

Environmental pollution is becoming a big challenge for the new developing and
developed world because of the industrial revolution, which on the one side is
facilitating the life but on the other side destroying the environment. The chemical,
textile, oil, and gas industries annually send billion of tons of hazardous effluents to
the groundwater and upper environment which is affecting badly the marine life,
human, and plant kingdom [1–3]. At present, there are a number of technologies that
have been discovered to cater the fast-growing problem; the most commonly
employed method of removing pollutants involves the use of various physical
techniques, chemical and biological processes, and thermal techniques to remove
commonly found organic and inorganic contaminants [4, 5]. Among many proposed
processes being developed for the decontamination of the organic-based pollutants,
biodegradation received the greatest attention. However, numerous organic
chemicals, especially those which are toxic or refractory, were found very resistant
to the biological degradation [6, 7]. Therefore, after the discovery of photocatalytic
splitting of water by Fujishima and Honda [8] in 1972, the researchers turned their
point of convergence toward semiconductor photocatalysis which proved very
effective in the degradation of even those pollutants which are highly difficult to
remove by the other means. In comparison with the traditional oxidation process, the
photocatalysis has significant advantages, for example, they can be used to degrade
and mineralize the dyes and chemicals completely to CO2 and H2O and are also
helpful in the degradation of very stable compounds which cannot be easily
degraded by the other processes, also they can work efficiently at ambient temper-
ature and pressure conditions, and they do not need any special supply of oxygen.
The other advantage of the photocatalysis is that it is a cheap process as compared to
the other oxidation process having no waste disposal problems [3]. A large number
of semiconductor photocatalysts have been investigated so far for the degradation of
different pollutants, such as ZnO [9, 10], WO3 [11, 12], Fe2O3 [13], CdSe [14], and

© Springer Nature Singapore Pte Ltd. 2018 1


J. Zhang et al., Photocatalysis, Lecture Notes in Chemistry 100,
https://doi.org/10.1007/978-981-13-2113-9_1
2 1 Mechanism of Photocatalysis

SrTiO3 [15]. In general, an ideal photocatalyst should have some basic properties,
such as they must be active under UV, visible light, or solar light, they should have
property of chemical and biological robustness, as well as they should be stable
toward photocorrosion. The other most important characteristic they must have is
that they should be nontoxic and must have low cost and easy availability
[16]. These all are key factors for a good photocatalyst.

1.2 TiO2 Photocatalysis

TiO2 is the most common catalyst used in the field of photocatalysis. From the
ancient times, TiO2 has been used as a white pigment. It can be extracted from the
ilmenite (FeTiO3) [17], perovskite (CaTiO3) [18], and titanite (CaTiSiO5) [19] ores.
TiO2 is mechanically, thermally, and chemically very stable. It has a very high
melting point of 1855  C and is insoluble in water, HNO3, HCl, and dilute H2SO4. It
can be dissolved in hot concentrated H2SO4 and HF [20]. The reports on TiO2 as
photocatalysis were found back in 1921 where Renz [21] from the University of
Lugano reported the partial reduction of titania during illumination under sunlight in
the presence of glycerol and in 1929 where Keidel [22] found that the dyes which are
quite stable to light and are not easily degradable are found to be decomposed when
adsorbed on lithopone or white lead in contact with TiO2 [23]. There they were
called TiO2 as photosensitizer. Titania is supposed to be an excellent photocatalyst
due to its enormous properties such as high activity, low cost, photostability,
nontoxicity, etc.

1.3 Mechanism of Photocatalytic Oxidation Reactions

In semiconductors the chemical reaction usually occurs due to the transfer of


electrons from the valance band to the conduction band as shown in the Fig. 1.1
As the number of orbitals (N) in the HOMO (highest occupied molecular orbital
(valance band)) and LUMO (lowest unoccupied molecular orbital (conduction
band)) increases, the energy to shift the electrons from the valance band to the
conduction band will decrease.
The detailed mechanism of the photocatalysis is quite a complex subject but the
basic principle is the same. Upon irradiation of the light with the energy equal or
more than the bandgap of the semiconductor photocatalyst, the electrons are excited
from the valance band of the catalyst to the conduction band, leaving the holes
behind in the valance band.
1.3 Mechanism of Photocatalytic Oxidation Reactions 3

Fig. 1.1 Changes in the


electronic structure of the
compound as the number of
atomic orbitals
(L) increases. (Reprinted
from Ref. [24], copyright
1997, with permission from
Elsevier)

TiO2 þ hν ! e þ hþ ð1:1Þ

These photo-generated electrons and holes can possibly recombine by releasing


the heat between 10 nanoseconds and 100 nanoseconds as mentioned by Hofmann
et al. by the measurement with the laser flash photolysis at room temperature
[3]. This recombination of the electrons and holes causes the low quantum efficiency
of the photocatalyst.

e þ hþ ! heat ð1:2Þ

The reports on the mechanism of the photocatalysis indicate that the recombina-
tion of photo-generated electrons and holes mainly occurs in the bulk of the catalyst
samples. This recombination process of the electrons and holes can be reduced
significantly if these charge-carrier species are separated by the addition of suitable
scavenger or incorporation of some of the trap sites on the surface as a result of
producing defects, surface adsorbents, or other sites.
If we provide sufficient time to the holes and electrons before they recombine,
then these produced charge carriers migrate to the surface of the catalyst and can
undergo the charge transfer to initiate the redox reactions with the pollutants
adsorbed on its surface which is shown in Fig. 1.2.
A valence band hole, h+, contains a strong oxidation power by having a redox
potential ranging from +1.0 to +3.5 V (measured vs normal hydrogen electrode
(NHE) at room temperature), depending on the semiconductor and pH. Thus, the
presence of the hole plays an important role in the photocatalytic degradation of the
pollutant present on the surface of the catalyst. The oxidation can take place either
due to the indirect oxidation via reaction with the surface-bound hydroxyl radical
4 1 Mechanism of Photocatalysis

Fig. 1.2 Mechanism for the photocatalytic activity on the surface of the semiconductor under the
irradiation of light. (Reprinted with permission from Ref. [25]. Copyright 2010 American Chemical
Society)

(•OH) (i.e., surface-trapped h+) or directly via h+ attached on the metal surfaces,
before it is trapped within the particle or at its surface.

hþ þ H2 Oads ! • OHads þ hþ ð1:3Þ


hþ þ HO ads ! • OHads ð1:4Þ
þ ðnþ1Þþ
h þM nþ
ads ! Mads ð1:5Þ

These produced surface hydroxyl radicals contain enormous power to oxide the
pollutants and hence mineralize them completely. Scheme 1.1 shows the formation
of different oxide species by the reactions of electrons and holes through various
pathways.
On the other hand, the electrons from the photoexcitation must also react to avoid
accumulation of excess charges within the catalyst particles. Therefore, the efficient
removal of the electrons can enhance the photocatalytic oxidation of the pollutants.
Oxygen is commonly employed as electron scavenger if the overall reaction target is
oxidation of dyes or organic compounds as it is very cost-effective and can be able to
dissolve in aqueous and other solutions. After the attack of the electron on the
oxygen, it can be reduced to the different oxygen activated species given by the
following equations:
1.4 Influence of Different Parameters on the Degradation of Pollutants 5

Scheme 1.1 Active oxygen species produced as a result of the irradiation of TiO2. (Reprinted with
permission from Ref. [3]. Copyright 1995 American Chemical Society)

O2 þ e ! O2 •  ð1:6Þ
O2 •  þ Hþ ! HO2 • ð1:7Þ

2 HO2 ! H2 O2 ð1:8Þ

H2 O2 þ hν ! OH ð1:9Þ
 • 
H2 O2 þ O2 ! OH þ OH þ O2 ð1:10Þ
H2 O2 þ e ! • OH þ OH ð1:11Þ

The generated hydroxyl radicals are the main oxidizing and reactive species with
the very short lifetime and having the oxidation potential of 2.8 V. Hence, these
produced hydroxyl radicals react with the substrate organic compounds and dye
molecules very rapidly to mineralize them completely in a very nonselective manner
as shown in Fig. 1.2.

1.4 Influence of Different Parameters on the Degradation


of Pollutants

The parameters like pH of the reaction solution, surface area of the catalyst,
concentration of the contaminant, catalytic loading, light wavelength, intensity,
etc. play an important role in the degradation of the pollutants on the surface of
6 1 Mechanism of Photocatalysis

TiO2. The changing in the different parameters can change the degradation results
which will be discussed one by one.

1.4.1 Catalyst Loading

The catalyst concentration for photoactivity in the sample is one of the important
parameters that control the degradation rate of the organic compounds. This is based
upon the initial concentration of the pollutant and volume of the solution being
examined. It has been generally noticed that the degradation rate of the dyes and
pollutants on the surface of the catalyst increases with an increase in the loading of
the catalyst [26]. The increased loading contributes to the high activity because of
the presence of more active catalyst sites at higher concentrations, which helps in the
formation of more hydroxyl species. Furthermore, when catalytic loading is
increased above the optimum value, there observed no substantial increase in the
degradation rate. This is due to the reason that the increased amounts of catalyst
particles in the solution make it turbid and thus block and scatter UV radiation for the
reaction to proceed, and therefore percentage degradation of the pollutant starts
decreasing [27–29]. At high concentrations of photocatalysts, the chances for parti-
cle agglomeration increase, which results in the abatement of the available surface
area for the light intake and hence reduces the generation of photoexcited electrons
and holes which affects the photoactivity and reaction rate drops significantly.
Table 1.1 shows the optimum catalyst concentration for the different dyes and
organic compound degradations with different light sources.

1.4.2 pH Effect

The pH of the reaction solution is another very important and determining factor for
the degradation of the pollutants in the presence of the catalyst. The degradation
reactions are immensely sensitive to the pH of photocatalytic reaction mixture. The
dependence of the reaction rate on pH is because of the reason that the pH controls
the adsorption of the organic compound onto the photocatalyst surface, and hence,
the degradation reactions rely on the ionization state and also on the surface charge
of the photocatalyst and the pollutant compound. The surface of titania can be
protonated or deprotonated under acidic or alkaline condition, respectively,
according to the following reactions [26, 51]:

TiOH þ Hþ ! TiOH2 þ ð1:12Þ


 
TiOH þ OH ! TiO þ H2 O ð1:13Þ
1.4 Influence of Different Parameters on the Degradation of Pollutants 7

Table 1.1 Effect of the catalytic loading on the degradation of pollutants [30–51]
Light TiO2 Optimum TiO2 concentration,
Pollutant type source Photocatalyst (g/L) g/L
Reactive Yellow 14 UV TiO2 1.0–6.0 4
BPA Solar TiO2 0–1.0 0.5
Chrysoidine R UV TiO2 0.5–1.0 5.0
Cibacron Yellow UV TiO2 0.1–4.0 4.0
LS-R
Phenol UV TiO2 0–4.0 2.0
Supra Blue BRL Visible K+-TiO2 0.25–2.0 1.5
Reactive Yellow 14 Solar TiO2 1.0–6.0 4.0
Acid Red 88 UV ZnO 2–12 12
Remazol Brilliant UV/solar TiO2 0.5–5.0 5.0
Blue R
Disperse Blue 1 UV TiO2 0.5–4.0 3.0
Acridine orange Visible ZnO 0–0.35 0.25
Amaranth UV TiO2 0.5–4.0 1.0
Bismarck UV TiO2 0.5–4.0 1.0
Acid Orange 8 UV TiO2 0.5–5.0 2.0
Acid Blue 45 UV TiO2 0.5–3.0 2.0
Xylenol UV TiO2 0.5–3.0 3.0
Acridine orange UV TiO2 0.5–3.0 2.0
Bromothymol UV TiO2 0.5–3.0 3.0
Fast green FCF UV TiO2 0.5–4.0 4.0
Reactive Orange 4 UV F-TiO2 1.0–5.0 4.0
Acid Blue 80 Solar TiO2 0.3–0.4 2.0
Rhodamine B Solar ZnO 0.05–0.4 0.3
Chromotrope 2B UV TiO2 0.5–5.0 5.0
Amido black 10B UV TiO2 0.5–5.0 5.0

Thus, as these reaction equations show, the surface of the catalyst will remain
positively charged in acidic medium and negatively charged in basic medium
[26]. Table 1.2 shows some pollutants and their optimum pH at which they show
the highest activity in terms of degradation/mineralization.

1.4.3 Surface Area and Morphology

The photocatalytic activity of semiconductor catalyst also decidedly confides on


surface and structural properties, such as surface area, crystal composition, distribu-
tion, porosity, bandgap, particle size, and surface hydroxyl group density [52]. Aver-
age crystal size in the heterogeneous catalysis is of primary importance because it is
directly related to its specific surface area which improves the efficiency of a catalyst.
Lin et al. [60] prepared the high-quality brookite TiO2 single-crystalline nanosheets
8 1 Mechanism of Photocatalysis

Table 1.2 Influence of pH on the photocatalytic degradation of pollutants [32, 36, 39–49, 53–59]
Pollutant type Light source Photocatalyst pH range Optimum pH
Phenol UV TiO2 4.1–12.7 7.4
Chrysoidine Y UV TiO2 3.0–9.0 9.0
m-Nitrophenol UV TiO2 4.1–12.7 8.9
Reactive Blue 4 UV Nd–ZnO 3.0–13.0 11
BPA Solar TiO2 2.0–10.0 6
Methylene Blue Visible La3+–TiO2 2–10 10
Supra Blue BRL Visible K+–TiO2 4.5–11.8 7.2
Reactive Orange 4 UV F–TiO2 1.0–9.0 3.0
Acid Red 88 Visible Ag–TiO2 0.2–1.8 1.8
Remazol Brilliant Blue R Solar TiO2 3.0–11.0 3.0
Disperse Blue 1 UV TiO2 3–11 3.0
Methyl orange UV Pt–TiO2 2.5–11.0 2.5
Amaranth UV TiO2 3.45–9.31 7.73
Bismarck UV TiO2 3.25–7.85 3.25
Acid Orange 8 UV TiO2 3.0–11.0 9.0
Acid Blue 45 UV TiO2 2.05–10.05 5.8
Acridine orange Visible ZnO 2.9–7.1 7.1
Bromothymol UV TiO2 2.2–9.0 4.35
Fast green FCF UV TiO2 3.0–11.0 4.4
Methyl Red UV Ag–TiO2 3.0–13.0 3–10.0
Acid Blue 80 Solar TiO2 2.0–10.0 10.0
Acid Red 29 UV TiO2 3.0–10.5 10.5

and tuned the catalyst activity from inert for the nanoflowers and nanospindles to the
highly active nanosheets by the control of the morphology of the samples. Zuo et al.
[61] prepared prism-shaped active faceted rutile Ti (III)-doped TiO2 with the control
of the morphology under acidic condition and hydrothermal treatment at 200  C for
12 h. They proved that facet effect is an important factor for heterogeneous
photocatalysts due to the surface atom arrangement and coordination which intrin-
sically determine the reactant adsorption on the surface of molecules, surface
distribution between photoexcited electrons and reactant molecules, as well as the
desorption of product molecules. Li et al. [62] prepared twist-like helix tungsten
nitrogen co-doped TiO2 for the enhanced degradation of phenol under visible light
irradiation and attributed the enhancement of the photoactivity to the large surface
area, special twist-like helix morphology, large pore size and better crystallinity of
the anatase.
1.4 Influence of Different Parameters on the Degradation of Pollutants 9

1.4.4 Effect of Temperature

Another factor that has a significant contribution in the photocatalytic degradation of


organic dyes and compounds is temperature. The temperature effect is not profound
if the reaction under observation has a little fluctuation; however, if the temperature
will be excessively higher or lower, it might change the course of degradation
reaction. Dissolved oxygen is one of the key elements that drive the photocatalysis,
as it helps in scavenging CB electrons and leads to the formation of hydroxyl
radicals. The higher or lower temperature causes the change in the percentage of
the dissolved oxygen in the samples and hence changes the degradation rate of the
reaction. Furthermore, the higher temperatures cause desorption of organic com-
pounds from the photocatalyst surface. Because of the fact that the photocatalytic
degradation of the compounds is a surface-based phenomenon, desorption of the
pollutant before the reaction with the electrons or holes due to the high temperature
results in a decrease of the reaction rate.

1.4.5 Effect of Contaminant Concentration

Another factor which can affect the degradation rate of the dyes and hazardous
chemicals is concentration in the reaction mixture. Most of the pollutants for the
degradation observe the pseudo-first-order kinetics which can be represented in
terms of Langmuir–Hinshelwood equation modified for the reactions occurring at
the liquid-solid interface:

ln ½C o =C ¼ kr Kt ¼ k 1 t ð1:14Þ

where k1 is apparent first-order rate constant, t is the time required for the initial
concentration of pollutant (Co) to reduce to (C), K is the equilibrium constant for the
adsorption of the pollutant on the surface of the catalyst, and Kr is the limiting rate of
the reaction [63].
In the case of the colored compounds like dyes, commonly the degradation rate
increases with the increase in the dye concentration, but after attaining a certain
critical concentration level, it starts decreasing. This decrease of the rate of degra-
dation with the increase of the concentration of pollutant can be attributed to the
screening of UV–visible light radiations by the dye molecules before reaching to the
surface of the catalyst. However, the concentration of the catalyst can be adjusted
based on the organic compound concentration, so that organic compound can be
considerably adsorbed on the photocatalyst surface and efficiently degraded. Most of
the degradation studies have employed a concentration of the organic compounds or
pollutant dyes in the range of 10–200 mg/L, which is similar with the concentration
of the common pollutants found in real wastewater [26]. Table 1.3 illustrates the
10 1 Mechanism of Photocatalysis

Table 1.3 Effect of contaminant concentration on the photoactivity of the catalysts [30, 39, 40, 42,
44–49, 51, 53, 54, 64–76]
Light Range of initial Optimum
Pollutant type source Photocatalyst concentration, mM concentration, mM
Phenol UV TiO2 0.13–0.71 0.13
m-Nitrophenol UV TiO2 0.13–0.71 0.13
Chrysoidine Y UV TiO2 0.13–1.0 0.75
Acid Orange 7 UV ZnO 0.003–0.009 0.003
Remazol Brilliant Solar TiO2 0.12–0.5 0.12
Blue R
Disperse Blue 1 UV TiO2 0.13–0.5 0.125
Amaranth UV TiO2 0.3–0.6 0.5
Bismarck UV TiO2 0.3–0.6 0.6
Reactive Orange UV F–TiO2 0.015–0.035 0.3
4
Acid Blue 45 UV TiO2 0.3–0.6 0.3
Xylenol Orange UV TiO2 0.3–0.6 0.5
Acridine orange UV TiO2 0.1–0.5 0.25
Ethidium UV TiO2 0.1–0.4 0.1
bromide
Bromothymol UV TiO2 0.15–0.5 0.35
Fast green FCF UV TiO2 0.031–0.125 0.031
Phenol UV Pr–TiO2 0.11–0.7 0.11
Acid Blue 80 Solar TiO2 0.03–0.2 0.03
Acid Red 18 UV ZnO 0.2–1.0 0.2
Chromotrope 2B UV TiO2 0.25–0.75 0.35
Amido black 10B UV TiO2 0.25–0.75 0.25

optimum dye concentration at which they can have maximum degradation under
solar and UV light.

1.4.6 Influence of Calcination Temperature

The calcination temperature is also an important factor which consummates the


betterment of the efficiency of the prepared photocatalyst by affecting the physical
and chemical properties of the catalyst. The calcination of the samples at high
temperatures can increase the photoactivity of the samples because the increased
temperature improves the crystallization of the sample and removes the unwanted
loosely bound impurities. Furthermore, the uncalcined samples contain the excess
number of water molecules on the surface which can reduce the surface active sites.
However, the reports showed that the increase in the calcination decreases the
surface area of the prepared catalyst because of the collapse of the pore at high
temperature [67]. Ohtani et al. [68] suggested the two main reasons for the increase
References 11

Table 1.4 Effect of calcination temperature on the degradation of pollutants [52, 67, 70–86]
Light Range of temperature Optimum temperature
Pollutant source Photocatalyst ( C) ( C)
Methyl orange UV Co2+–TiO2 150–700 200
Azo fuchsine Visible TiO2 300–800 400
Methylene blue UV Ce4+–TiO2 300–500 400
Methyl orange UV N–TiO2 0–400 300
Rhodamine B UV La3+–TiO2 600–900 700–800
Methylene blue UV Sm3+–TiO2 500–800 600
4-Chlorophenol UV N–TiO2 300–700 500
o-Cresol Visible Pt–TiO2 150–750 150
Methylene blue UV– WO3–TiO2 450–850 550
visible
Acid orange 52 UV/solar Zn2+–TiO2 400–550 450
BPA UV Zr–TiO2 500–700 700
Methylene blue Visible N–TiO2 400–700 500
Methyl orange UV– Co–TiO2 500–850 650
visible
Phenol UV– I–TiO2 400–600 400
visible
Rhodamine B Visible F–TiO2 120–180 180
Phenol UV WOX–TiO2 400–800 700
Methyl orange Solar Fe3+–TiO2 200–600 300
2,4,6- UV TiO2 350–650 550
Tribromophenol

in the photoactivity of the samples: one is the high surface area of the prepared
catalyst which gives the high room for the substrate dyes to adsorb on its surface and
the other is crystallinity of the samples which allows less recombination of the
photoexcited electrons and holes and increases the photocatalytic activity. Therefore,
the increase in the calcination temperature increases the crystallinity but on the other
side decreases the specific surface area of the samples. So, these two requirements
are used to be partially fulfilled at a moderate calcination temperature [69]. Table 1.4
illustrates some catalysts with the optimum calcination temperature for the degrada-
tion of different types of pollutants.

References

1. Arakawa H, Aresta M, Armor JN et al (2001) Catalysis research of relevance to carbon


management: progress, challenges, and opportunities. Chem Rev 101(4):953–996
2. Esswein AJ, Nocera DG (2007) Hydrogen production by molecular photocatalysis. Chem Rev
107(10):4022–4047
3. Hoffmann MR, Martin ST, Choi W et al (1995) Environmental applications of semiconductor
photocatalysis. Chem Rev 95(1):69–96
12 1 Mechanism of Photocatalysis

4. Hashim MA, Mukhopadhyay S, Sahu JN et al (2011) Remediation technologies for heavy metal
contaminated groundwater. J Environ Manag 92(10):2355–2388
5. Haji-Saeid SM, Sampa MH, Safrany A et al (2012) Radiation processing techniques in
remediation of pollutants, and the role of the IAEA in supporting capacity building in devel-
oping countries. Radiat Phys Chem 81(8):1040–1044
6. Khan FI, Husain T, Hejazi R (2004) An overview and analysis of site remediation technologies.
J Environ Manag 71(2):95–122
7. Yeung AT, Gu YY (2011) A review on techniques to enhance electrochemical remediation of
contaminated soils. J Hazard Mater 195(0):11–29
8. Fujishima A, Honda K (1972) Electrochemical photolysis of water at a semiconductor elec-
trode. Nature 238(0):37–38
9. Ba-Abbad MM, Kadhum AAH, Mohamad AB et al (2013) Visible light photocatalytic activity
of Fe3+-doped ZnO nanoparticle prepared via sol–gel technique. Chemosphere 91
(11):1604–1611
10. Burton P, Peterson E, Boyle T et al (2010) Synthesis of high surface area ZnO(0001) plates as
novel oxide supports for heterogeneous catalysts. Catal Lett 139(1–2):26–32
11. Bignozzi CA, Caramori S, Cristino V et al (2013) Nanostructured photoelectrodes based on
WO3: applications to photooxidation of aqueous electrolytes. Chem Soc Rev 42(6):2228–2246
12. Tian L, Ye L, Liu J et al (2012) Solvothermal synthesis of CNTs–WO3 hybrid nanostructures
with high photocatalytic activity under visible light. Catal Commun 17(0):99–103
13. Franking R, Li L, Lukowski MA et al (2013) Facile post-growth doping of nanostructured
hematite photoanodes for enhanced photoelectrochemical water oxidation. Energy Environ Sci
6(2):500–512
14. Bang JU, Lee SJ, Jang JS et al (2012) Geometric effect of single or double metal-tipped CdSe
nanorods on photocatalytic H2 generation. J Phys Chem Lett 3(24):3781–3785
15. Wang J, Yin S, Zhang Q et al (2003) Mechanochemical synthesis of fluorine-doped SrTiO3 and
its photo-oxidation properties. Chem Lett 32(6):540–541
16. Bhatkhande DS, Pangarkar VG, Beenackers AACM (2002) Photocatalytic degradation for
environmental applications – a review. J Chem Technol Biotechnol 77(1):102–116
17. Ilmenite (2015) http://en.wikipedia.org/wiki/Ilmenite
18. Perovskite (2017) http://en.wikipedia.org/wiki/Perovskite
19. Titanite (2017) http://en.wikipedia.org/wiki/Titanite
20. Budavari S (1996) An encyclopedia of chemicals, drugs, and biologicals, merck index, 12th
edn. Merck, Whitehouse Station, p 1617
21. Renz C (1921) Photoreactions of oxides of titanium, cerium and earth acids. Helv Chim Acta 4
(0):961–968
22. Keidel E (1929) Die Beeinflussung der Lichtechtheit von Teerfarblacken durch Titanweiss (The
effect on the light fastness of coal tar dye paints by titanium white). Farben-Zeitung 34
(0):1242–1243
23. Serpone N, Emeline AV, Horikoshi S et al (2012) On the genesis of heterogeneous
photocatalysis: a brief historical perspective in the period 1910 to the mid-1980s. Photochem
Photobiol Sci 11(7):1121–1150
24. Mills A, Hunte SL (1997) An overview of semiconductor photocatalysis. J Photochem
Photobiol A Chem 108(1):1–35
25. Chen X, Shen S, Guo L et al (2010) Semiconductor-based photocatalytic hydrogen generation.
Chem Rev 110(11):6503–6570
26. Vinu R, Madras G (2011) Photocatalytic degradation of water pollutants using Nano-TiO2. In:
Zang L (ed) Energy efficiency and renewable energy through nanotechnology. Springer,
London, pp 625–677
27. Rauf M, Ashraf SS (2009) Fundamental principles and application of heterogeneous
photocatalytic degradation of dyes in solution. Chem Eng J 151(1):10–18
References 13

28. Wang CC, Lee CK, Lyu MD et al (2008) Photocatalytic degradation of C.I. Basic Violet
10 using TiO2 catalysts supported by Y zeolite: an investigation of the effects of operational
parameters. Dyes Pigments 76(3):817–824
29. Macedo LC, Zaia DAM, Moore GJ et al (2007) Degradation of leather dye on TiO2: a study of
applied experimental parameters on photoelectrocatalysis. J Photochem Photobiol A Chem 185
(1):86–93
30. Ahmed S, Rasul MG, Martens W et al (2011) Advances in heterogeneous photocatalytic
degradation of phenols and dyes in wastewater: a review. Water Air Soil Pollut 215(1–4):3–29
31. Muruganandham M, Swaminathan M (2006) Photocatalytic decolourisation and degradation of
Reactive Orange 4 by TiO2-UV process. Dyes Pigments 68(2):133–142
32. Kaneco S, Rahman MA, Suzuki T (2004) Optimization of solar photocatalytic degradation
conditions of bisphenol A in water using titanium dioxide. J Photochem Photobiol A Chem 163
(3):419–424
33. Qamar M, Saquib M, Muneer M et al (2005) Titanium dioxide mediated photocatalytic
degradation of two selected azo dye derivatives, chrysoidine R and acid red 29 (chromotrope
2R), in aqueous suspensions. Desalination 186(1–3):255–271
34. Kositzi M, Poulios I, Samara K et al (2007) Photocatalytic oxidation of Cibacron Yellow LS-R.
J Hazard Mater 146(3):680–685
35. Chiou CH, Wu CY, Juang RS (2008) Influence of operating parameters on photocatalytic
degradation of phenol in UV/TiO2 process. Chem Eng J 139(2):322–329
36. Chen LC, Huang CM, Tsai FR (2007) Characterization and photocatalytic activity of K+-doped
TiO2 photocatalysts. J Mol Catal A Chem 265(1–2):133–140
37. Muruganandham M, Shobana N, Swaminathan M (2006) Optimization of solar photocatalytic
degradation conditions of Reactive Yellow 14 azo dye in aqueous TiO2. J Mol Catal A Chem
246(1–2):154–161
38. Behnajady MA, Moghaddam SG, Modirshahla N (2009) Investigation of the effect of heat
attachment method parameters at photocatalytic activity of immobilized ZnO nanoparticles on
glass plate. Desalination 249(3):1371–1376
39. Saquib M, Muneer M (2002) Semiconductor mediated photocatalysed degradation of an
anthraquinone dye, Remazol Brilliant Blue R under sunlight and artificial light source. Dyes
Pigments 53(3):237–249
40. Saquib M, Abu Tariq M, Haque MM (2008) Photocatalytic degradation of disperse blue 1 using
UV/TiO2/H2O2 process. J Environ Manag 88(2):300–306
41. Pare B, Jonnalagadda SB, Tomar H (2008) ZnO assisted photocatalytic degradation of acridine
orange in aqueous solution using visible irradiation. Desalination 232(1–3):80–90
42. Abu Tariq M, Faisal M, Muneer M (2005) Semiconductor-mediated photocatalysed degrada-
tion of two selected azo dye derivatives, amaranth and bismarck brown in aqueous suspension. J
Hazard Mater 127(1–3):172–179
43. Saquib M, Muneer M (2003) Titanium dioxide mediated photocatalyzed degradation of a textile
dye derivative, acid orange 8, in aqueous suspensions. Desalination 155(3):255–263
44. Abu Tariq M, Faisal M, Saquib M et al (2008) Heterogeneous photocatalytic degradation of an
anthraquinone and a triphenylmethane dye derivative in aqueous suspensions of semiconductor.
Dyes Pigments 76(2):358–365
45. Faisal M, Abu Tariq M, Muneer M (2007) Photocatalysed degradation of two selected dyes in
UV-irradiated aqueous suspensions of titania. Dyes Pigments 72(2):233–239
46. Haque MM, Muneer M (2007) TiO2-mediated photocatalytic degradation of a textile dye
derivative, bromothymol blue, in aqueous suspensions. Dyes Pigments 75(2):443–448
47. Saquib M, Abu Tariq M, Faisal M (2008) Photocatalytic degradation of two selected dye
derivatives in aqueous suspensions of titanium dioxide. Desalination 219(1–3):301–311
48. Vijayabalan A, Selvam K, Velmurugan R (2009) Photocatalytic activity of surface fluorinated
TiO2-P25 in the degradation of Reactive Orange 4. J Hazard Mater 172(2–3):914–921
49. Su Y, Deng L, Zhang N et al (2009) Photocatalytic degradation of C.I. Acid Blue 80 in aqueous
suspensions of titanium dioxide under sunlight. React Kinet Catal Lett 98(2):227–240
14 1 Mechanism of Photocatalysis

50. Byrappa K, Subramani AK, Ananda S et al (2006) Photocatalytic degradation of rhodamine B


dye using hydrothermally synthesized ZnO. Bull Mater Sci 29(5):433–438
51. Qamar M, Saquib M, Muneer M (2005) Photocatalytic degradation of two selected dye
derivatives, chromotrope 2B and amido black 10B, in aqueous suspensions of titanium dioxide.
Dyes Pigments 65(1):1–9
52. Ahmed S (2011) Impact of operating conditions and recent developments in heterogeneous
photocatalytic water purification process. Crit Rev Environ Sci Technol 42(6):601–675
53. Chiou CH, Wu CY, Juang RS (2008) Photocatalytic degradation of phenol and m-nitrophenol
using irradiated TiO2 in aqueous solutions. Sep Purif Technol 62(3):559–564
54. Qamar M, Saquib M, Muneer M (2005) Semiconductor-mediated photocatalytic degradation of
anazo dye, chrysoidine Y in aqueous suspensions. Desalination 171(2):185–193
55. Zhou Y, Lu SX, Xu WG (2009) Photocatalytic activity of Nd-doped ZnO for the degradation of
C.I. Reactive Blue 4 in aqueous suspension. Environ Prog Sustain Energy 28(2):226–233
56. Parida KM, Sahu N (2008) Visible light induced photocatalytic activity of rare earth titania
nanocomposites. J Mol Catal A Chem 287(1–2):151–158
57. Anandan S, Kumar PS, Pugazhenthiran N et al (2008) Effect of loaded silver nanoparticles on
TiO2 for photocatalytic degradation of Acid Red 88. Sol Energy Mater Sol Cells 92(8):929–937
58. Huang M, Xu C, Wu Z et al (2008) Photocatalytic discolorization of methyl orange solution by
Pt modified TiO2 loaded on natural zeolite. Dyes Pigments 77(2):327–334
59. Gupta AK, Pal A, Sahoo C (2006) Photocatalytic degradation of a mixture of Crystal Violet
(Basic Violet 3) and Methyl Red dye in aqueous suspensions using Ag+ doped TiO2. Dyes
Pigments 69(3):224–232
60. Lin H, Li L, Zhao M (2012) Synthesis of high-quality brookite TiO2 single-crystalline
nanosheets with specific facets exposed: tuning catalysts from inert to highly reactive. J Am
Chem Soc 134(20):8328–8331
61. Zuo F, Bozhilov K, Dillon RJ (2012) Active facets on titanium(III)-doped TiO2: an effective
strategy to improve the visible-light photocatalytic activity. Angew Chem 124(25):6327–6330
62. Li J, Xu J, Dai WL et al (2008) One-pot synthesis of twist-like helix tungsten-nitrogen-codoped
titania photocatalysts with highly improved visible light activity in the abatement of phenol.
Appl Catal B Environ 82(3):233–243
63. Boarini P, Carassiti V, Maldotti A et al (1998) Photocatalytic oxygenation of cyclohexane on
titanium dioxide suspensions: effect of the solvent and of oxygen. Langmuir 14(8):2080–2085
64. Daneshvar N, Rasoulifard MH, Khataee AR (2007) Removal of C.I. Acid Orange 7 from
aqueous solution by UV irradiation in the presence of ZnO nanopowder. J Hazard Mater 143
(1–2):95–101
65. Sobana N, Swaminathan M (2007) The effect of operational parameters on the photocatalytic
degradation of acid red 18 by ZnO. Sep Purif Technol 56(1):101–107
66. Chiou CH, Juang RS (2007) Photocatalytic degradation of phenol in aqueous solutions by
Pr-doped TiO2 nanoparticles. J Hazard Mater 149(1):1–7
67. An T, Liu J, Li G et al (2008) Structural and photocatalytic degradation characteristics of
hydrothermally treated mesoporous TiO2. Appl Catal A Gen 350(2):237–243
68. Ohtani B, Ogawa Y, Nishimoto SI (1997) Photocatalytic activity of amorphousanatase
mixture of titanium(IV) oxide particles suspended in aqueous solutions. J Phys Chem B 101
(19):3746–3752
69. Gupta S, Tripathi M (2011) A review of TiO2 nanoparticles. Chin Sci Bull 56(16):1639–1657
70. Hua DS, Kunfeng Z, Daorong L et al (2009) Synthesis, characterization and photocatalytic
activity of Co2+ -doped titania. J Phys Conf Ser 188(1):012–018
71. Wang J, Zhao G, Zhang Z et al (2007) Investigation on degradation of azo fuchsine using visible
light in the presence of heat-treated anatase TiO2 powder. Dyes Pigments 75(2):335–343
72. Du Y, Du M, Qiao Y et al (2007) Ce(IV) doped TiO2 thin films: characterization and
photocatalysis. Colloid J 69(6):695–699
73. Chen Y, Zhang S, Yu Y et al (2008) Synthesis, characterization, and photocatalytic activity of
N-doped TiO2 nanotubes. J Dispers Sci Technol 29(2):245–249
References 15

74. Quan X, Tan H, Zhao Q et al (2007) Preparation of lanthanum-doped TiO2 photocatalysts by


coprecipitation. J Mater Sci 42(15):6287–6296
75. Xiao Q, Si Z, Yu Z et al (2008) Characterization and photocatalytic activity of Sm3+-doped
TiO2 nanocrystalline prepared by low temperature combustion method. J Alloys Compd 450
(1–2):426–431
76. Yu H, Zheng X, Yin Z et al (2007) Preparation of nitrogen-doped TiO2 nanoparticle catalyst and
its catalytic activity under visible light. Chin J Chem Eng 15(6):802–807
77. Chen HW, Ku Y, Kuo YL (2007) Effect of Pt/TiO2 characteristics on temporal behavior of
o-cresol decomposition by visible light-induced photocatalysis. Water Res 41(10):2069–2078
78. Saepurahman AMA, Chong FK (2010) Dual-effects of adsorption and photodegradation of
methylene blue by tungsten-loaded titanium dioxide. Chem Eng J 158(3):418–425
79. Chen C, Wang Z, Ruan S (2008) Photocatalytic degradation of C.I. Acid Orange 52 in the
presence of Zn-doped TiO2 prepared by a stearic acid gel method. Dyes Pigments 77
(1):204–209
80. Gao B, Lim TM, Subagio DP et al (2010) Zr-doped TiO2 for enhanced photocatalytic degra-
dation of bisphenol A. Appl Catal A Gen 375(1):107–115
81. Sathish M, Viswanathan B, Viswanath RP (2007) Characterization and photocatalytic activity
of N-doped TiO2 prepared by thermal decomposition of Ti–melamine complex. Appl Catal B
Environ 74(3–4):307–312
82. Sun CG, Tao L, Fan ML et al (2009) Replication route synthesis of mesoporous titanium–cobalt
oxides and their photocatalytic activity in the degradation of methyl orange. Catal Lett 129
(1–2):26–38
83. Hong X, Wang Z, Cai W et al (2005) Visible-light-activated nanoparticle photocatalyst of
iodine-doped titanium dioxide. Chem Mater 17(6):1548–1552
84. Ren G, Gao Y, Liu X et al (2010) Synthesis of high-activity F-doped TiO2 photocatalyst via a
simple one-step hydrothermal process. Reac Kinet Mech Catal 100(2):487–497
85. Piszcz M, Tryba B, Grzmil B (2009) Photocatalytic removal of phenol under UV irradiation on
WOx –TiO2 prepared by sol–gel method. Catal Lett 128(1–2):190–196
86. Liu J, Qin W, Zuo S (2009) Solvothermal-induced phase transition and visible photocatalytic
activity of nitrogen-doped titania. J Hazard Mater 163(1):273–278

You might also like