0% found this document useful (0 votes)
28 views14 pages

1 s2.0 S1389556712000457 Main

Uploaded by

padminisuthakar
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
28 views14 pages

1 s2.0 S1389556712000457 Main

Uploaded by

padminisuthakar
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 14

Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276

Contents lists available at SciVerse ScienceDirect

Journal of Photochemistry and Photobiology C:


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

Invited review

Kinetics and mechanisms of charge transfer processes in photocatalytic systems:


A review
Liwu Zhang a , Hanan H. Mohamed b , Ralf Dillert a , Detlef Bahnemann a,∗
a
Institute of Technical Chemistry, Leibniz University of Hannover, Callinstr. 3, D-30167 Hannover, Germany
b
Physical Chemistry Department, Helwan University, Ain Helwan, Cairo, Egypt

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

Article history: Charge carrier transfer processes are very important and play a vital role in photocatalytic reactions. A
Received 30 May 2012 fundamental understanding of the kinetics and mechanisms of these charge transfer processes is cru-
Received in revised form 10 July 2012 cial from the viewpoint of developing efficient photocatalysis systems for large-scale industrialization.
Accepted 23 July 2012
In this work, recent efforts concerning the understanding of the kinetics and the mechanisms of the
Available online 31 July 2012
charge transfer in photocatalytic processes have been reviewed. Fundamental aspects involved in these
charge transfer processes, such as charge generation, charge trapping, charge recombination, and elec-
Keywords:
tron and hole transfer are primarily discussed. Moreover, some recent studies focusing on enhancing the
Photocatalysis
Charge transfer
photocatalytic efficiency by improving the charge transfer and separation are also reviewed.
Kinetics © 2012 Elsevier B.V. All rights reserved.
Mechanism
Charge recombination

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
2. Charge carrier generation and trapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
2.1. Light absorption and charge carrier generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
2.2. Charge carrier trapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
3. Charge carrier recombination kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
4. Charge carrier transfer kinetics and mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
4.1. Electron transfer reactions involving transition metal ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
4.2. Multiple electron transfer reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
4.3. Hole transfer kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
5. Improvement of the charge carrier transfer process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
5.1. Noble metal loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
5.2. Plasmonic structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
5.3. Graphenic carbon coatings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274

1. Introduction have been proven to be able to effectively degrade a great variety


of pollutants or to even decompose water into H2 and O2 . Over the
Semiconductor particles employed as photocatalysts have past decade, TiO2 photocatalysis has been extensively studied and
found great potential in applications such as environmental reme- the commercialization of the process has been achieved in many
diation and solar energy conversion [1–8]. Many photocatalysts fields [1–8]. The photocatalytic process is a consequence of the
interaction between the semiconductor particles activated under
light irradiation with the surrounding substances. Electron–hole
pairs are generated upon light absorption in these activated semi-
∗ Corresponding author.
conductor particles, where they can then recombine or participate
E-mail addresses: [email protected],
[email protected] (D. Bahnemann).
in reductive and oxidative reactions, respectively. The reductive

1389-5567/$20.00 © 2012 Elsevier B.V. All rights reserved.


http://dx.doi.org/10.1016/j.jphotochemrev.2012.07.002
264 L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276

and oxidative capacity of the photogenerated electrons and holes is the penetration of light in the semiconductor. The absorption
are determined by the conduction band and valence band potential threshold wavelength of a semiconductor is decided by the band
of the respective semiconductor particle. The photogenerated holes gap energy via
in TiO2 show, for example, very high oxidative activity and there-
1240
fore TiO2 has been extensively investigated concerning its role as a = (3)
Eg
photocatalyst for the light-induced degradation of environmental
contaminants. It has been shown in extensive degradation stud- It has been reported that the effective band gap energy is dif-
ies employing many different model compounds that most of the ferent between bulk solid particles and nanoparticles due to the
organic contaminants present in water and even in air can be fully quantum size effect. TiO2 nanoparticles with size of 3 nm show an
decomposed [6,9,10]. apparent band gap increase of 0.25 eV [13]. It was reported that the
Though its ability to achieve the total mineralization of most lower edge of the conduction band of a semiconductor is given by
pollutants render photocatalysis into a very attractive oxidation [14]
technology, research has yet to succeed in developing a sound and
h2 1.8e2
comprehensive description and understanding of all the involved E(R) = Eg + ∗ 2
− (4)
8m0 m R εR
phenomena, with the challenge reaching its maximum mani-
festation when real wastewater and varying solar illumination where m0 is the electron mass in vacuum and m* is the reduced
intensities become the implicated issues. In a photocatalytic reac- effective mass of the semiconductor. Accordingly, when a semicon-
tion, charge carrier transfer has been widely acclaimed to be ductor has a reduced effective mass significantly smaller than the
very important and plays a vital role. Therefore, the fundamen- free electron mass, a large variation of its band gap can be expected.
tal study of the dynamics of the charge transfer processes is very For some semiconductors with larger effective mass, the quantiza-
important from the viewpoints of developing novel photocatalytic tion only occurs when the particle sizes reach the molecular level.
systems and a large-scale industrialization of the photocatalytic The increase in band gap will reduce the light absorption of the
processes. During the recent years, extensive studies have been car- semiconductor, but the electrons and holes may possess higher
ried out focusing on the kinetics and the mechanistic details of the reduction and oxidation power, respectively, due to the shift of the
charge transfer processes involved in a photocatalytic reaction, for band edges. The band gap energy is also closely related to the crys-
the purpose of a better understanding of the underlying mecha- tal phase. For example, TiO2 is mainly found in three crystal forms,
nism of photocatalysis. Because the charge transfer processes are i.e., anatase, rutile and brookite, the band gap energies of which
usually very fast, the laser photolysis and the pulse radiolysis tech- are found to be 3.2 eV, 3.0 eV and 3.3 eV, respectively. During the
nique are commonly employed to investigate the charge transfer past decade, considerable efforts have been dedicated to reducing
kinetics. the band gap energy of TiO2 for the purpose of visible light uti-
Although photocatalytic processes and their applications have lization. A frequently employed mechanism is to introduce new
been widely reviewed during the past decades [1,3–5,8,11], and energy states within the band gap or to change the band edge by
Majima and coworkers reviewed the underlying one-electron oxi- forming more dispersive conduction or valence bands by doping.
dation processes using time resolved spectroscopy [12], there are Doping with anions has attracted much attention since N-doped
hardly any reviews focusing on the kinetics and the mechanisms of TiO2 was the first material reported to exhibit photocatalytic activ-
charge transfer processes in photocatalytic systems. Therefore, the ity under visible light irradiation [15]. Following this work, sulfur-
present review focusses on kinetics studies, charge transfer mech- and carbon-doped TiO2 have also been reported to show visible-
anisms and some approaches for improving the charge transfer of light photocatalytic activity [16–21]. It has also been found that
photocatalytic systems. the doping of transition metals into TiO2 can extend its absorption
edge into the visible light region [22–25].
2. Charge carrier generation and trapping Theoretically, a photon the energy of which is higher than the
bandgap energy of TiO2 can generate shallow valence band holes
2.1. Light absorption and charge carrier generation and “hot” electrons in the conduction band well above its lower
edge. These “hot” electrons should, in principle, exhibit more reduc-
The first step during heterogeneous photocatalysis is the gen- tive ability due to their higher potential energy. However, in reality,
eration of electron–hole pairs in the employed semiconductor it is very difficult to use these more active charge carriers for pho-
particles. Under irradiation, a semiconductor can absorb photons tocatalytic processes, because the charge carrier thermalization
with an energy h greater or equal to its bandgap energy Eg (i.e., is very fast. Turner et al. [26] monitored the photoconductivity
3.2 eV for anatase TiO2 and 3.0 eV for rutile TiO2 ), resulting in the of conduction band (CB) electrons injected into Evonik-Degussa
formation of electron–hole pairs (Eq. (1)). Aeroxide P 25 (frequently employed as a standard photocatalyst
powder) from an adsorbed dye by using time-resolved terahertz
TiO2 + h → TiO2 (e−
cb
+ h+
vb ) (1) spectroscopy. By measuring the frequency-dependent complex
conductivity on a subpicosecond time scale, these authors have
where e− cb
represents a conduction band (CB) electron and h+vb shown that carrier cooling within the conduction band occurs in
represents a positive hole in the valance band (VB) of the semicon- about 300 fs.
ductor, respectively. After their generation according to reaction Several groups also studied the hole thermalization in the
(1), the fate of the generated electron and hole can follow several valence band. Grela et al. [28] found that photons absorbed by
pathways, including bulk recombination, surface recombination, nanocrystalline anatase TiO2 particles at 254 nm are found to be 7.7
migration to the semiconductor surface and capture by (adsorbed) times more efficient than those at 366 nm for driving the photocat-
substrate molecules. alytic oxidation of salicylate. The same research group showed in a
Within the semiconductor, the intensity of light follows the quantitative photochemical study of the photocatalytic oxidation
exponential law of 3-nitrophenol in polycrystalline and metastable TiO2 nanoparti-
I = I0 exp(−˛l) (2) cles that quantum yields sensibly increase with photon energy on
both materials [29]. These works show that both, hole transfer and
where I is the light intensity inside the semiconductor, I0 is the hole trapping can preceed hole thermalization in TiO2 , which is also
initial light intensity, ˛ is the reciprocal absorption length and l supported by some other studies [27,30,31]. Sawada and coworkers
L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276 265

for holes resulting from band-to-band excitation at either 266 or


355 nm.
It is found that the lifetimes of the trapped charge carriers can
be very long, while the actual trapping process is very fast. Many
groups have investigated the lifetimes of the trapped electrons
[43]. Most results show that the electron trapping lifetime is longer
than hundreds of picoseconds. Some studies even reported that the
lifetime of the trapped electron can be on a timescale of months.
Kuznetsov et al. [44] studied the light-induced charge separation
and storage in anatase titanium oxide gels. They observed that
under prolonged UV-laser irradiation, more than 14% of Ti4+ cen-
ters can be converted into Ti3+ . The latter’s lifetime can be as long
as several months in the absence of oxygen. The trapped electrons
are responsible for a “dark” absorption continuum in the spectral
range from 350 nm to 2.5 mm. The reported lifetime of the trapped
Fig. 1. Typical TRG signals from the TiO2 /SCN− (aq) interfaces with various pump electrons varies over a wide timescale, which is closely related to
beam powers. The intensities of the pump beam power are (a) 4; (b) 5; (c) 7; (d) the presence or absence of hole or electron scavengers (such as
8 mW. molecular oxygen). Shen et al. [45] studied the lifetimes of trapped
Ref. [27] Copyright 1999 ACS. holes in rutile and anatase TiO2 . The lifetimes of holes in rutile were
found to be longer than hundreds of ps, which is longer than the
hole lifetime in anatase TiO2 .
[27] applied the femtosecond transient reflecting grating (TRG) The electron trapping energy states in TiO2 have been eval-
method to solid/liquid interfaces to successfully observe the pho- uated by several research groups. It has been reported that the
toinduced ultrafast relaxation dynamics at TiO2 (0 0 1)/KSCN (aq) trapping states in nanocrystalline TiO2 are located at energy levels
interfaces (Fig. 1). Their experimental value for the charge car- about 0.2–0.9 eV below the conduction-band edge [46–48]. Hupp
rier relaxation time of 690 fs at the (1 0 0) surface is much faster and coworkers used time-resolved photoacoustic spectroscopy
than all previously estimated or reported values, which suggests (TRPAS) to probe directly the trap state energetics of photogener-
that the observed hole transfer process is a new type transfer ated electrons in nanocrystalline colloidal anatase titanium dioxide
process beyond the electron transfer according to the Marcus the- dispersed in aqueous solution. The electron trapping sites were
ory. The authors attribute the new-type process to nonequilibrium determined to lie, on average, 0.8 eV below the conduction band
hole transfer before thermalization of the photoexcited energy. edge [34]. Serpone et al. performed photoluminescence measure-
This ultrafast nonequilibrium hole transfer suggests that not only ments on colloidal anatase TiO2 particles with different size, and
equilibrated holes can be transferred across an interface, but also shallow trap levels were ascertained at 0.41 and 0.64 eV (2.1 nm
non-thermalized holes. particle), 0.51 eV (13.3 nm), and 0.53 eV (26.7 nm) below the con-
duction band edge of anatase [49]. Rutile TiO2 shows almost
the same photoluminescence spectrum from shallow traps as the
2.2. Charge carrier trapping anatase particles, indicating the same origin of shallow traps, i.e.,
the presence of oxygen vacancies in rutile as well as anatase par-
Following their generation, the charge carriers can migrate to ticles [5]. The energetics of trapped holes have been evaluated
the surface where they are trapped in the subsurface and sur- indirectly. Serpone and coworkers [50] used the pulse radiolysis
face states of the particle. These trapped charge carriers can be technique to study the reaction of 13-nm TiO2 particles with • OH
detected and investigated because of their strong transient optical radicals. The product of the • OH reaction with the particles was
absorptions. It has been shown in early laser flash photolysis studies identified as a trapped hole at the particle surface, the redox level
employing titanium dioxide anatase nanoparticles that the trapped of which was found to lie approximately 1.3 eV above the valence
electron exhibits a strong optical absorption around 650 nm while band.
the trapped hole can be characterized by a transient absorption pre- Many studies have examined the trapping sites of the photo-
dominantly at shorter wavelengths around 430 nm or even shorter generated holes and electrons. Generally, it is proposed that Ti4+
[32,33]. Time-resolved laser flash photolysis measurements in col- cations at the surface of the titanium dioxide particle can be trap-
loidal titanium dioxide suspensions show that trapping is a very fast ping sites for electrons [36,52–54]. There are also reports that the
process. Many studies show that trapping of an electron occurs on electrons can be trapped in the bulk [53,55,56]. Hoffmann et al. [57]
the sub-picosecond time scale [34–39]. For example, Rothenberger performed laser flash photolysis measurements on TiO2 and pro-
et al. performed picosecond and nanosecond transient absorption posed that the CB electrons are trapped at two different Ti3+ sites
experiments on titanium dioxide and observed that the electron (Eqs. (5) and (6))
trapping time was faster than 30 ps, i.e., the time resolution of their
laser system [40]. Serpone et al. performed a picosecond laser flash e− cb + Ti4+ -OH ↔ Ti3+ -OH (5)
photolysis study on anatase titanium dioxide colloid solutions and e−
+ Ti 4+
(bulk) → Ti 3+
(bulk) (6)
cb
observed that the absorption spectra of trapped electrons as well
as of trapped holes are fully developed after a laser pulse of 30 ps where Ti3+ -OH is a surface-trapped electron and Ti3+ (bulk) denotes
[41]. Another laser flash photolysis study by Bowman and cowork- an electron trapped in the bulk. The dynamic equilibrium of Eq.
ers indicated that approximately 200 fs was needed for the full (5) represents a reversible trapping of a CB electron in a shallow
development of the transient absorption spectrum of the electron trap. Eq. (6) represents the irreversible trapping of an electron in
[38]. Hole trapping on the surface of TiO2 occurs as fast as elec- a deep trap. Serpone et al. [38] and Bowman et al. [41] have also
tron trapping. Yang and Tamai [42] observed that the hole trapping reported the existence of two different traps from the analysis of the
timescale at the surface of colloidal anatase TiO2 particles is about charge carrier recombination kinetics in titanium dioxide colloidal
50 fs by detecting its transient absorption signal at 520 nm. Tamaki solutions and in dispersions. Further studies show that photoex-
et al. [31,35] estimated a hole trapping timescale of about 200 fs cited electrons are preferentially trapped at surface OH groups on
266 L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276

Fig. 3. Comparison of the time-resolved transient absorption of 0.7 g/L Q(quantum


Fig. 2. EELS spectra from TMA-covered TiO2 (1 1 0), oxidized prior to TMAA adsorp-
sized)-TiO2 solution and the time-resolved diffuse reflectance of the dry Q-powder.
tion. (a) TMA-covered surface prior to UV; (b–e) after irradiation with a 100 W Hg
The smooth solid lines are fits to second order kinetics with a baseline.
arc lamp; (f) exposure of e to 240 L of O2 . Inset: correlation between the isobutene
and CO2 photodesorption yields and the Ti3+ signal generated during photolysis. Ref. [38] Copyright 1995 ACS.
Ref. [51] Copyright 2003 ACS.

In summary, the initial steps of a photocatalytic reaction are


charge carrier generation and trapping. It is important to under-
TiO2 . Henderson et al. [51] tried to identify surface sites associated
stand how the charge carriers are generated and trapped. The
with charge transfer and trapping during the photodecomposition
charge generation and “hot” charge carriers have been discussed.
of the trimethyl acetate acid (TMAA) on the rutile (1 1 0) surface,
The charge trapping is a very fast process (picosecond time scale),
employing scanning tunneling microscopy (STM), electron energy
while the lifetimes of trapped charge carriers can be very long.
loss spectroscopy (EELS), and photodesorption. This research team
The energy states of the electron traps in TiO2 are reported to be
observed that UV irradiation of the TMAA-covered surface in the
located about 0.2–0.9 eV below the conduction band edge. Con-
absence of O2 resulted in a feature at about 0.8 eV that grew with
troversies still exist concerning the trapping sites in TiO2 for the
additional UV irradiation (spectra c–e), as shown in Fig. 2. This
photogenerated electrons and holes.
feature is indicative of Ti3+ resulting from photogenerated elec-
trons trapped at the surface of TiO2 (1 1 0) as Ti3+ -OH groups [51].
Additional evidence from Fourier transform infrared analyses and 3. Charge carrier recombination kinetics
luminescence measurements also confirmed the surface OH groups
as preferred trapping sites [53,58]. Since most of the photogenerated charge carriers will recom-
There are still controversies regarding the trapping sites in bine following their generation, it is important to understand the
TiO2 for photoexcited holes. Some reports propose that the charge carrier recombination kinetics in order to improve the
holes are trapped at the titanium dioxide surface in adsorbed overall photoconversion efficiency. Photogenerated charge carri-
hydroxyl groups yielding weakly adsorbed hydroxyl radicals Ti4+ - ers can recombine in a radiative or in a nonradiative way. The
• OH [50,59]. Howe and Grätzel found that the hole is trapped at recombination process has been examined during laser flash pho-
a subsurface oxygen anion based on their ESR investigations [60]. tolysis studies from the rather rapid depletion of the transient
Other groups also presented evidence for subsurface hole trapping absorption spectra recorded, as seen in Fig. 3 [38]. The nonradia-
[52,61,62]. Electrostatic calculations regarding hole trapping at the tive way is believed to dominate the recombination event in TiO2
rutile TiO2 (1 1 0) surface suggest that holes prefer the trapping sites [31,35,55,65], while there is increasing interest in the investigation
in the near-surface region rather than in the bulk or at surface sites of the radiative recombination of photogenerated charge carriers
[63]. Micic et al. [64] show that the EPR technique can be employed in TiO2 [49,66,67].
to characterize hole traps in aqueous TiO2 colloids. Their results are The recombination kinetics of the charge carriers have been
consistent with the conclusions of Howe and Grätzel that the hole studied in detail by many research groups [38,40,41]. Grätzel and
adduct is a subsurface oxygen anion radical. Bahnemann and co- coworkers [40] performed for the first time picosecond time-
workers proposed an extended model with two different types of resolved spectroscopy to elucidate the dynamics of the charge
trapping sites being more readily suited in explaining their exper- carrier reactions within a TiO2 particle. They monitored the dynam-
imental observations [52]. They assume that at least two different ics of charge carrier trapping and recombination in colloidal
trapping sites for holes exist on the surface of a TiO2 particle. Holes particles of anatase TiO2 (diameter 12 nm) in aqueous solu-
trapped in energetically deep traps can be characterized by their tion excited by picosecond or nanosecond laser photolysis, and
transient absorption around 430 nm, while the holes initially resid- interpreted the results by a stochastic kinetic model. While the
ing in shallow traps do not show such spectral features. The holes absorption spectrum of the trapped electron appears within the
trapped in shallow traps can be excited thermally into the valence leading edge of the 30-ps laser pulse, the trapping of the hole
band, achieving equilibrium with free holes. It is proposed that is found to be a much slower process requiring 250 ns. The EPFL
deeply trapped holes are chemically equivalent to surface-bound research group observed that at high electron–hole pair concen-
hydroxyl radicals. Weakly trapped holes being readily detrapped tration in the particles, their recombination follows a second order
apparently possess an electrochemical potential close to that of law, with the rate coefficient being (3.2 ± 1.4) × 10−11 cm3 s−1 .
free holes and can therefore be considered to be chemically similar Intraparticle recombination becomes first order at very low charge
to the latter. These shallow traps are probably created by surface carrier occupancy, the mean lifetime of a single electron/hole pair
imperfections of the semiconductor nanocrystal [52]. was determined to be 30 ± 15 ns [40]. Bowman and co-workers
L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276 267

studied the subpicosecond dynamics of anatase titanium diox-


ide sols with particle sizes of about 2 nm [38], Fig. 3 shows the
time-resolved transient absorption of 0.7 g/L Q(quantum sized)-
TiO2 solution and the time-resolved diffuse reflectance of the dry
Q-powder. The Bowman research team found that the average
lifetime of an electron/hole pair is 23 ± 5 ps, and that substantial
electron/hole recombination occurs within the first 30 ps. A second-
order recombination rate constant of (1.8 ± 0.7) × 10−10 cm3 s−1
for trapped electrons with holes has been reported in another
study of the same research group [68]. However, there are other
reports expressing considerable disagreement with the second-
order recombination kinetics [31,69]. Colussi and coworkers
reported stochastic calculations in compact 2-D lattices modeling
the competition between electron–hole recombination and reac-
tive processes on the surface of illuminated anatase TiO2 colloidal
particles [69]. They showed that the decay of single and multiple
carrier pairs displays nonclassical kinetic behavior, and proposed
that bimolecular carrier recombination never follows second-
order kinetics: single excitons decay exponentially and multiple Fig. 4. Transient absorption study of nc-TiO2 films in anaerobic environment and in
the presence of gas-phase ethanol in concentrations ranging from 0.3% to 2.9%. Data
pairs annihilate with second-order rate coefficients asymptotically
in the absence of ethanol (0%) are also shown. Data were collected at an excitation
approaching a t−1/2 dependence. wavelength of 337 nm and a probe wavelength of 800 nm. O.D. is the transient
Serpone and coworkers have studied the effects of particle size absorption change.
on the charge carrier recombination dynamics in semiconductor Ref. [54] Copyright 2006 ACS.
particles [41]. Colloidal anatase TiO2 sols with mean particle diam-
eters of 2.1, 13.3, and 26.7 nm, respectively, were examined by
charge recombination occurs by means of tunneling and with-
picosecond transient absorption and emission spectroscopy. The
out trapping of charges. The charge recombination rate is found
authors observed that the smaller the particle, the greater the frac-
to be sensitive around room temperature for dye-sensitized TiO2
tion of electron/hole pairs that recombine within 20 ps. Absorption
films, which is attributed to a reduced barrier height facilitating
decay for the 2.1 nm particles is a simple first-order process, and
direct charge recombination. This research team proposed that
electron/hole recombination is 100% complete within 10 ns. For the
slow charge recombination is not caused by charge traps but is
13.3 and 26.7 nm specimens the transient absorption decay follows
instead caused by a barrier between a positive charge and an elec-
distinct second-order biphasic kinetics; the decay times of the fast
tron.
components decrease with increasing particle size. Within 10 ns
The charge recombination can be effected by trapping and the
after the laser pulse, about 90% or more of the photogenerated
interfacial charge transfer. If a hole trapping agent such as an alco-
electron/hole pairs have recombined. The faster components are
hol is present, the holes will be effectively trapped, resulting in
explained by the recombination of shallow-trapped charge carriers,
less recombination. The recombination kinetics can be different
whereas the slower components ( > 20 ns) reflect the recombina-
under vacuum or in the presence of molecular oxygen. Durrant
tion of deep-trapped electrons and holes.
and co-workers have studied the photochemical reduction of oxy-
It has been shown by the early laser-pulsed transient absorption
gen adsorbed on nanocrystalline TiO2 films by means of transient
spectroscopic studies on TiO2 particles that the recombination of
absorption spectroscopy [54]. They observed that the transient
the photogenerated electron/hole pairs is a very fast process (occur-
absorption signal in the absence of ethanol (Fig. 4, 0%), which
ing within a time scale of ns). However, several research groups
exhibited a decay with a half-life of about 25 ␮s under anaerobic
proposed a longer timescale for the charge carrier recombination
conditions, decays with a half-time of about 12 ␮s for an oxygen
[70,71]. Yamakata et al. [70] observed long-lived electrons surviv-
concentration of 21%, indicative of oxygen reduction competing
ing in the sub-second region in their study of the time-resolved
directly with the electron/hole recombination reaction observed
infrared absorption spectroscopy of photogenerated electrons in
in the absence of hole scavengers. Ishibashi used the time-resolved
platinized TiO2 particles. They propose that the highly sensitive
infrared absorption spectra to investigate the water-induced and
detection of the IR absorption made it possible to trace the kinetics
oxygen-induced decay kinetics of free electrons in TiO2 [73]. They
of a small number of the electrons until a delay time as long as 1 s. To
found that the decay kinetics of the electrons were sensitive to the
achieve conditions close to a real photocatalytic system, Katoh and
presence of O2 or H2 O. On bare TiO2 , O2 from the gas-phase cap-
coworkers [71] used a low-intensity pulsed laser to excite a TiO2
tured the electrons at a delay time of 10–100 ␮s. In the presence of
film ensuring that less than one electron/hole pair was generated
water vapor, holes reacted with hydroxyls group within 2 ␮s, and
in each particle per pulse. They observed that the half-life for elec-
the recombination decay of the electrons was obstructed thereafter
tron/hole recombination is ca. 1 ␮s for a nanocrystalline anatase
[73].
TiO2 film dipped in N2 -saturated deaerated water. Durrant and
In summary, most of the reports show that the charge recom-
coworkers also observed a transient absorption signal exhibiting
bination follows a second order law. The recombination of the
a decay with a half-time of 25 ␮s in their study on nanocrystalline
photogenerated electron/hole pairs is a fast process in a time scale
titanium dioxide films [54]. They attributed this long decay time to
of ns, while some groups reported longer lifetimes, even as long
the low excitation intensity, while the electron/hole recombination
as 1 s. The charge recombination rate can be affected by trapping,
will take place on much faster time scales at high excitation inten-
interfacial charge transfer and temperature.
sities probably due to trap filling resulting from the high excitation
densities.
Katoh and Furube have studied the effect of temperature on 4. Charge carrier transfer kinetics and mechanism
the charge recombination rate in bare and dye-sensitized TiO2
films [72]. They observed that the charge recombination rate is Exciting a semiconductor with any light source causes interband
not sensitive to temperature for bare TiO2 films, indicating that transitions, excitonic transitions or below-band gap transitions. If
268 L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276

The dynamics of interfacial charge transfer have been studied


since the early 1980s [76,77]. When either the hole or the electron
are trapped by a surface adsorbed catalyst or by an appropriate
acceptor, the non- trapped electrons or holes might react with suit-
able acceptor molecules in solution. The overall reaction will then
be composed of two steps:

1. Formation of an encounter complex of the electron (or hole)


acceptor with the semiconductor particle. The rate of this process
will most likely be diffusion limited.
2. Interfacial electron transfer being an electrochemical step that
Fig. 5. Schematic illustration of spatial and energetic distribution of electron traps
involves a Faradic current across the semiconductor-solution
in TiO2 films. interface and that is characterized by a rate parameter (cm s−1 ).
Reprinted with permission from Ref. [1] (a). Copyright 2007 RSC.
A detailed kinetic treatment of the above reaction steps has been
carried out by Grätzel et al. [6] who derived the following equation:
the photon energy is larger than the band gap energy, interband
1 1
1 R

transitions dominate, exciting electrons from the valence band into
= + (7)
the conduction band in the fs timescale. Tamaki et al. [31,35] have kobs 4R2 kct D
proposed that both, electrons and holes, move randomly to the where R is the sum of the radii of the semiconductor particle and
surface of the photocatalyst where they are trapped in the sub- the electron (or hole) and D is the sum of their respective diffusion
picosecond time domain. Electrons and holes can also be trapped coefficients, kobs is the familiar rate constant obtained experimen-
at bulk trapping sites. Tamaki et al. [31,35] suggested that electrons tally and kct is the electrochemical rate parameter. There are two
are capable of migrating between the surface and the bulk traps that limiting cases of Eq. (7):
are in equilibrium since these species are energetically equivalent
and probably very close to the conduction band edge. Fig. 5 shows (a) Heterogeneous charge transfer being rate determining and
the energy levels for the surface trapped electrons and for the bulk much slower than diffusion (kct  D/R). In this case (Eq. (7))
electrons suggested from surface relaxation dynamics. At the sur- reduces to:
face of the photocatalyst the trapped electron and the trapped hole
can recombine or they can be transferred to acceptor (A) or donor kobs = 4R2 kct (8)
(D) molecules, respectively thus reacting with them. Following this (b) Heterogeneous charge transfer being faster than diffusion,
charge transfer back reactions between the reduced acceptor and which controls the overall reaction rate (kct  D/R). In this case
the hole or between the oxidized donor and the electron can occur, the well known Smoluchowski expression is obtained from (Eq.
in particular, when the species are strongly adsorbed to the surface (7)):
[1,54,74,75]. The mechanistic steps of photocatalysis on a semi-
conductor nanoparticle are summarized in Fig. 6. According to this kobs = 4DR (9)
mechanism the interfacial charge transfer efficiency is limited by
two important processes: the competition between charge carrier Diffusion effects can be eliminated by adsorption or chemi-
recombination and trapping followed by the competition between cal fixation of the acceptor at the semiconductor particle surface
trapped carrier recombination and interfacial charge transfer. resulting in the following relation:
The recombination of electrons and holes is an important elec- ı
tronic process in semiconductors. Rectification, photoconductivity, kct = (10)
ct
and transistor behavior are critically dependent on the life time of
where  ct is the average time required for the charge carrier to
the injected mobile carriers. The recombination rate also plays a
tunnel across the interface and ı is the reaction layer thickness.
critical role in photocatalytic reactions with semiconductor parti-
cles.
4.1. Electron transfer reactions involving transition metal ions

The photocatalytic reduction of transition metal ions is an


important process because of two effects: (1) the transformation
of the ions to often less toxic species and (2) the deposition of the
reduced metals on the semiconductor catalyst surface, e.g., for the
recovery of expensive and useful metals. Moreover, noble metal
islands deposited on semiconductor surfaces frequently exhibit
unusual catalytic and optoelectronic properties [78]. It is well
known that at the nanometer scale, the optical, electronic and
catalytic properties of transition-metal nanoclusters are highly
sensitive to their size and shape. Whereas deposition of metal
nanoparticles improves both charge separation as well as inter-
facial charge transfer kinetics, the mechanism responsible for such
an improvement is yet to be fully understood on the nanoscale.
One main reason for this paucity of mechanistic information is the
lack of the experimental methods able to follow these processes
in real time [79]. Fundamental understanding of the kinetics and
mechanisms involved in the reduction of metal nanoclusters on
Fig. 6. Schematic representation of the general mechanistic steps occuring in het- nanometer-scale semiconductors is important both from the fun-
erogeneous photocatalysis on a semiconductor nanoparticle. damental and from the practical point of view. It is believed that
L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276 269

the main catalytic properties including selectivity, activity, lifetime


and stability [80,81] depend on the catalyst size [82], its surface
composition and its structure [83], which in turn require greater
control over the catalyst syntheses. The reduction of the metal ions
is based on their reaction with the photogenerated electrons in the
photocatalytic system:

M n+ + TiO2 (xe−
CB
) → M (n−x)+ (11)

It has been reported that Cu2+ can be reduced to Cu0 by TiO2


electrons [83,84]. Herrmann et al. [77], however, suggested that in
the presence of TiO2 suspensions Cu2+ ions are not reduced to Cu0
but to Cu+ even in the absence of molecular oxygen.
Recently, Bahnemann and coworkers [85–87] studied the reduc-
tion of Ag+ , Cu2+ , Zn2+ , Mn2+ , and Au3+ employing stored electrons
on anatase TiO2 nanoparticles. Ag+ , Cu2+ , and Au3+ ions were
reduced through single-, two- and three-electron transfer reac- Fig. 7. The possible multi-electron transfer processes of the heme/TiO2 (e− ) system
with organohalide pollutants.
tions, respectively, to form Ag0 , Cu0 , and Au0 nanoparticles,
Reprinted with permission from Ref. [89]. Copyright 2003 ACS.
respectively, which have been detected by their typical surface
plasmon resonance bands. These authors also found that neither
the reduction of Zn2+ nor that of Mn2+ can be induced using stored Obare et al. [89,90] studied multi-electron transfer reactions
TiO2 electrons in these systems, attributing it to the thermody- of organohalide pollutants at heme-functionalized nanocrystalline
namic infeasibility of the respective reduction processes. anatase TiO2 . Using this catalyst resulted in the formation of sta-
A multistep reduction of noble metal ions (Ag+ [86] and Au3+ ble carbene products in a greater than 60% yield. In this system a
[87]) on the surface of TiO2 followed by the transfer of excess elec- synergy between the molecular catalysts and the semiconductor
trons to the deposited metal particles has been proposed according particles was observed. The mechanism of organohalide reduction
to the following equations: mediated by Heme functionalized TiO2 systems is illustrated in
Fig. 7 and described by the following equations [91]:
xe−
TiO
x+
+ Mads → M0 (Reduction) (12)
2
Heme Heme
+ RX → + RX• − Slow (20)
nM 0 + Mn0 (Nucleation) (13) TiO2 (ne− ) TiO2 (ne− )
+ +
Mn0 + Mads → Mn+1 (Autocatalytic growth) (14) Heme Heme
→ Fast (21)
+
TiO2 (ne− ) TiO2 ((n − 1)e− )
Mn+1 + e−
TiO2
→ −
Mn+1 (15)
Heme Heme
e− , M+
+ RX → + RX• − (22)
TiO2 TiO2 TiO2
0
Mn+1 −→ Mx0 (16)
Gao et al. [92] have studied the reactions of excess electrons in
Mx0 + My0 → Mx+y
0
(Coalescence) (17) TiO2 produced by radiolysis employing steady state and pulse radi-
olysis techniques. The authors analyzed the kinetics of the electron
0
Mx+y + Ze− 0
→ (Mx+y )
Z−
(Transfer of electrons to Ag particles) transfer reactions with several scavengers including NO2 − , NO3 − ,
TiO 2 O2 , and H2 O2 using different TiO2 nanocrystal sizes. The authors
(18) found that the rates of the electron reactions depend on the respec-
tive particle size. Several scavengers including ClO2 − , ClO3 − , NO2 − ,
and NO3 − induce a decay of the TiO2 electron absorption signal
z− (z−2)−
0
(Mx+y ) + 2H+
ads
0
→ (Mx+y ) + H2 (H2 generation) (19) predominantly by single pseudo-first order processes. The rate of
reaction of the above ions in the large nanocrystallite systems is
4.2. Multiple electron transfer reactions found to be 2–10 times faster than in the respective small particle
systems. Gao et al also found that the rate of the reduction of the
Upon illumination by high light intensities, charges can accu- different scavengers tends to increase with the driving force.
mulate at the semiconductor–electrolyte interface. Depending on Recently, Mohamed et al. [85] studied the multi-electron
the surface pre-treatment, the presence of surface states and transfer reactions induced by stored electrons in anatase TiO2
the kinetics facilitating the interfacial electron transfer, the accu- nanoparticles employing the stopped flow technique. They investi-
mulated charges may sometimes be used to induce multiple gated the multi-electron reduction of common electron acceptors
electron transfer to or from adsorbed substrates. This charge accu- such as O2 , H2 O2 , and NO3 − . The authors observed that the rate
mulation is critical for the redox reactions of simple inorganic of the electron absorbance decay depends on the concentration
molecules such as water oxidation, oxygen reduction, nitrogen of these electron acceptors (cf. Fig. 8) [85]. They found that the
reduction, nitrate reduction, chlorinated hydrocarbons reduction rate constant of the first initial decay (kI obs = 2.0 × 104 M−1 s−1 ) and
or CO2 reduction. Multi-electron transfer reactions are important that of the second slower decay (kII obs = 5.7 × 103 M−1 s−1 ) for the
processes which avoid high-energy free radical intermediates and reaction of the stored TiO2 electrons with molecular oxygen were
can yield desired reaction products under mild conditions. Multi- similar to the values of the rate constants of the two slower pro-
electron storage and hydrogen generation by UV(A)-illumination cesses kIII (7.3 ± 1.5 × 103 M−1 s−1 ) and kIV (1.0 ± 0.2 × 103 M−1 s−1 )
was achieved in aqueous dispersions of ultrafine TiO2 particles as reported by Gao et al. [92] who studied the reaction of electrons
when the amphiphilic viologen derivative N-tetradecyl-N -methyl- formed by radiation chemical processes in small TiO2 particles and
4,4 -dipyridinium dichloride (C14 MV2+ ) was used as an electron who suggested that these slow processes are due to the reaction of
relay [88]. Two-electron reduction of C14 MV2+ was coupled with excess TiO2 electrons with molecular oxygen in the bulk.
H2 generation in alkaline medium in the presence of Pt acting as Moreover, the multi-electron reduction of nitrate has recently
catalyst when being co-deposited onto the TiO2 particles. received special attention in view of pollution control [93–96].
270 L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276

[37] have probed the charge carrier dynamics of opaque, aque-


ous suspensions of Degussa P-25 TiO2 employing femtosecond
time-resolved diffuse reflectance spectroscopy. They studied the
interfacial hole transfer dynamics of the P-25 TiO2 /SCN− complex
as a function of the thiocyanate ion concentration. It was found
that a dramatic increase in the population of trapped charge car-
riers is established within the first few picoseconds. To explain
this observation, it is proposed that the interfacial charge trans-
fer of an electron from the SCN− to a hole on the photoexcited
TiO2 effectively competes with the electron–hole recombination
on an ultrafast time scale [37]. Bahnemann et al. have studied
the reactivity of free and trapped holes for the photocatalytic
oxidation of the model compounds dichloroacetate, DCA− , and
thiocyanate, SCN− , by employing time-resolved laser flash pho-
tolysis [52]. Detailed spectroscopic investigations of the processes
occurring upon band gap irradiation in the colloidal aqueous
anatase TiO2 suspensions (24 Å) in the absence of any hole scav-
Fig. 8. Time profiles of the decay of the e− absorbance at 600 nm ([e− ] = 5.7 × engers showed that while electrons are trapped instantaneously,
TiO
2 TiO
2
10−4 M, 6 e− /particle) observed upon mixing of their aqueous suspension with
i.e., within the duration of the laser flash (20 ns), at least two
TiO 2
different concentration of dissolved O2 in aqueous solutions at pH 2.3 (HCl). The
different types of traps have to be considered for the remain-
solid lines represent the respective double exponential fits. ing holes. The authors propose that deeply trapped holes, h+ tr , are
Reprinted with permission from Ref. [85]. Copyright 2011 ACS. rather long-lived and unreactive, i.e., they are transferred neither
to DCA− nor to SCN− ions. Shallowly trapped holes, h+ tr∗ , on the
other hand, are in a thermally activated equilibrium with free
Mohamed et al. [85] have found also that nitrate ions are reduced
holes thus exhibiting a very high oxidation potential. The over-
to ammonia through an eight electron transfer including the inter-
all yield of trapped holes can be considerably increased when
mediate formation of molecular nitrogen according to the following
small platinum islands are present on the TiO2 surface which
equations:
act as efficient electron scavengers competing with the unde-
2NO3 − + 12H+ + 10e− → N2 + 6H2 O Eredox sired e− /h+ recombination. It is proposed that while molecular
TiO 2
oxygen, O2 , reacts in a relatively slow process with trapped
= 1.25 V (vs. NHE) (23) electrons (k2 = 7.6 × 107 L mol−1 s−1 ), the adsorption of the model
compounds DCA− and SCN− on the TiO2 surface prior to the band
gap excitation appears to be a prerequisite for an efficient hole
N2 + 4e− + 5H+ → N2 H5 + Eredox = −0.23 V (vs. NHE) (24) scavenging. In the case of DCA− the detailed kinetic analysis of the
TiO 2
time-resolved spectroscopic data reveals an extremely good cor-
N2 H5 + + 2e−
TiO
+ 3H+ → 2NH4 + Eredox = 1.27 V (vs. NHE) (25) relation with independent adsorption measurements. Moreover,
2
calculations using the Marcus electron transfer theory for adiabatic
On the other hand, Gao et al. [92] have observed the one- instead of processes which result in a reorientation energy  = 0.64 eV suggest
the multi-electron reduction of nitrate ions by excess TiO2 electrons that also in the case of SCN− the hole transfer occurs in the adsorbed
using pulse radiolysis technique. state [52]. Other studies show that strongly absorbed ionic species
In recent publications [92,97–99], the reaction of electrons in primarily react with free holes while weakly adsorbed molecules
TiO2 with H+ and water in the presence of noble metal coatings mainly react with long-lived h+ tr in a diffusion controlled process
have been investigated employing the pulse radiolysis technique. [32,33].
Kasarevic-Popovic et al. [97] found that the reaction of TiO2 elec- Tamai and coworkers [42] studied the hole transfer by in situ
trons with H+ /H2 O is a pseudo first order reaction depending on monitoring of carrier dynamics in anatase TiO2 nanoparticles with
both the Pt0 and the H+ concentration. It was shown that the energy femtosecond laser spectroscopy. They assigned the broad absorp-
of the adsorbed hydrogen atoms on the Pt0 surface is not suffi- tion band around 520 nm observed immediately after band-gap
cient to abstract hydrogen from 2-propanol. A similar observation excitation for the system without SCN− to shallowly trapped
was reported for Au/TiO2 by Hussein and Rudham [98]. Behar and holes. The absorption from the trapped holes at 520 nm cannot be
Rabani [99] studied the kinetics of hydrogen production upon the observed in the presence of SCN− , which is attributed to the ultra-
reaction of TiO2 electrons with H+ and water in the presence of dif- fast interfacial hole transfer between TiO2 nanoparticles and SCN− .
ferent surface metal catalysts (Pd0 , Pt0 and Au0 ). They found that The hole and electron trapping times were estimated to be <50 and
when the TiO2 nanoparticles are partially coated with Au0 instead 260 fs, respectively. The rate of the hole transfer from nanosized
of Pd0 or Pt0 , a higher than expected molecular hydrogen level is TiO2 colloid to SCN− is found to be comparable to that of the hole
observed and attributed this to a short chain reaction involving trapping. [42]
hydrogen abstraction from 2-propanol. Sawada and coworkers [27] applied a femtosecond transient
reflecting grating (TRG) method to observe the photoinduced ultra-
4.3. Hole transfer kinetics fast relaxation dynamics at TiO2 (0 0 1)/KSCN (aq) interfaces. For
these TiO2 /SCN− interfaces, the estimated charge transfer time by
It has been proposed in many studies that the photocatalytic Marcus theory is 1.8 ns, and Bahnemann et al. [52] reported that the
oxidation of organic compounds may occur by either oxidation time constant for the hole transfer process there is 5 ␮s. The exper-
via a surface-bound hydroxyl radical (indirect oxidation), which imental value of 110–690 fs reported by Sawada et al, however, is
is formed by a hole trapped at the particle surface, or via the much faster than the estimated and reported values. They propose
valence band hole before it is trapped (direct oxidation). Interfa- that the observed hole transfer process is a new-type transfer pro-
cial hole transfer involved in the TiO2 photocatalysis process has cess beyond the limits defined by the Marcus theory. They offered
been studied by several groups [37,52]. Bowman and coworkers a possible explanation for their observation that this new-type
L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276 271

process is a nonequilibrium hole transfer before thermalization


of the photoexcited energy. It is thus proposed that the nonequi-
librium holes need to be taken into account during interfacial
reactions [27].
Both holes and hydroxyl radicals are acting as oxidizing species
which has been observed in a study of the photocatalytic degra-
dation of acetate [100]. Both formate and formaldehyde were
detected as the only products of the photocatalytic degradation
of acetate in acidic suspensions (pH 3.0), while glycolate and for-
mate were detected as the main products in alkaline suspensions
(pH 10.6) [100]. Hence, both holes and hydroxyl radicals are pro-
posed to contribute to the photocatalytic oxidation of acetate by
comparing the product distribution with that obtained in homoge-
neous solutions upon oxidation of acetate with hydroxyl radicals
or by direct one-electron oxidation on a Pt electrode [101]. The
intermediate formation of hydroxyl radicals as the main oxidizing
species in photocatalytic reactions is also supported by photo-
catalytic degradation experiments of aromatic compounds in the
presence of TiO2 . It has been observed that the intermediates dur- Fig. 9. Transient decay observed at 800 nm following 308-nm laser excitation of
ing the photocatalytic degradation of aromatic compounds are deaerated 5.5 mM TiO2 solution containing (a) no Au, (b) 126 ␮M Au (8 nm), and (c)
126 ␮M Au (5 nm).
typically hydroxylated structures, which is consistent with the
Ref. [126] Copyright 2005 Elsevier.
intermediates found in the reaction of similar aromatics with a
known source of hydroxyl radicals. Moreover, different ESR stud-
ies have also confirmed the existence of hydroxyl radicals in TiO2 absorption decay at 675 nm following 308-nm laser pulse exci-
suspensions under irradiation [59,102]. However, later ESR stud- tation of a deaerated TiO2 solution (Fig. 9) [126]. They observed
ies by Micic et al. did not support the intermediate formation a transient absorption (cf trace a in Fig. 9), indicating that elec-
of hydroxyl radicals [64], and recently Bui et al. reported that tron trapping is completed within the laser pulse duration of 10 ns.
hydroxyl radical are not involved in the oxidation mechanism of The charge distribution between the gold and the photoexcited
benzene on TiO2 [103]. Imanishi and coworkers [104] proposed TiO2 nanoparticles is apparently already completed within the
a new mechanism for the surface water photooxidation reaction laser pulse duration, based on the absorption-time profiles shown
on TiO2 . They suggest that the water photooxidation reaction is in Fig. 9. It is proposed that the decay of the absorbance in the
not initiated by an electron transfer oxidation of Ti-OH to form microsecond time scale represents the fraction of electrons that are
hydroxyl radicals but is rather initiated by a nucleophilic attack lost in the recombination process. Kamat et al. also performed laser
of an H2 O molecule (Lewis base) to a surface-trapped hole (Lewis flash photolysis experiments of TiO2 colloids in the presence of 5-
acid). and 8-nm Au nanoparticles, and observed a decrease in the maxi-
mum absorbance immediately following the laser pulse (cf traces
b and c in Fig. 9). They propose that the decrease in the maximum
5. Improvement of the charge carrier transfer process
absorbance can be regarded as an indication that fewer electrons
remain in the TiO2 particles. The pronounced difference in the max-
Due to a much longer charge carrier transfer time as compared
imum absorbance observed in the absence and in the presence of
with the electron-hole recombination time, most of the photo-
Au colloids is explained by a higher number of electrons that reside
generated charge carriers recombine instead of being captured.
in the Au nanoparticles. It is also observed in this study that more
Therefore, it is not surprising that the photocatalytic efficiency is
electrons can be stored on 5-nm Au nanoparticles than on 8-nm
usually found to be very low. To improve the efficiency, it is there-
Au nanoparticles. Close examination of the initial decay fraction
fore important to enhance the charge separation and to suppress
reveals that the charge recombination is significantly decreased in
the charge recombination. During the past decades, many meth-
the presence of the gold nanoparticles. The role of Au colloids acting
ods have been developed to improve the charge carrier transfer as
as electron sinks able to store and shuttle photogenerated elec-
well as their separation in a photocatalytic system. Herein, we will
trons is also indicated by a decreased band-edge emission intensity
mainly review the methods of improving charge carrier transfer
in the ZnO/Au system, which has been observed in a picosecond
and separation by noble metal loading, plasmonic structures, and
time-resolved transient fluorescence study [127].
conjugated graphenic carbon coatings.
Iwata et al. investigated femtosecond time-resolved near-
infrared spectra of TiO2 and Pt/TiO2 powders to follow trapping,
5.1. Noble metal loading detrapping and lifetimes of excited electrons [129]. They found the
relaxation and trapping timescales to be about the same for Pt/TiO2
Nobel metal coating of TiO2 is one of the most efficient strate- and for TiO2 , while the recombination timescale was much longer
gies to improve the efficiency of photocatalysts. It has been for the Pt/TiO2 system. When Pt was loaded to TiO2 , an additional
widely reported that surface deposition of Pt, Au, Ag, and Pd on decaying process of 2.3 ps was observed. The authors proposed this
TiO2 can significantly enhance the photocatalytic activity of TiO2 decay component to represent the transfer of generated electrons
[105–115,73,116–125]. Although there are still some controversies from TiO2 to Pt, which is consistent with the known increase of the
concerning the underlying enhancement mechanism, it is widely overall catalytic activity induced by the Pt cocatalyst.
agreed that noble metal particles on the TiO2 surface can act as elec- Berr et al. have studied the delayed photoelectron transfer in
tron sinks, storing and shuttling photogenerated electrons, thereby Pt-decorated CdS nanorods under hydrogen generation conditions
facilitating the charge carrier separation and thus reducing the [128]. They observed that under hydrogen generation conditions
charge carrier recombination rate. (i.e., in presence of the hole-scavenger sodium sulfite), the rate
Kamat and coworkers have studied the charge transfer between of the photoelectron transfer to the Pt islands is slowed down
anatase TiO2 and Au colloids by monitoring the transient by an order of magnitude, as shown in Fig. 10. They explain this
272 L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276

Fig. 10. (A) Experimental transient absorption traces of 10 wt% Pt–CdS without additives and in the presence of either hole- or electron-scavenger. (B) Schematic represen-
tation of the hypothesized explanation.
Ref. [128] Copyright 2012 Wiley.

surprising phenomenon by localization and delocalization of the photocatalyzed decolouration of dyes for rutile TiO2 , for anatase
electron wavefunction in the presence and absence of holes on PC500, and for the anatase-rutile mixture (75%–25%) of Degussa
the nanostructure, respectively (Fig. 10B). Different degrees of P25 [132]. The metal loading resulted in the largest enhancement
delocalization lead to different overlaps of the electronic wavefunc- of the UV photocatalyzed dye degradation, when Pt was deposited
tion with the Pt clusters and the trap states, which consequently on rutile. On P25, Pt significantly reduced the efficiency of the UV
changes the trapping and transfer rates. The authors propose that photocatalyzed decolouration of both dyes. It was inferred that
charge transfer rate optimizations alone are not sufficient for platinum influences the UV photocatalysis by facilitating the elec-
improving the performance of any photocatalytic nanosystems. tron transfer to adsorbed oxygen, and that this facilitation is much
Instead, the collective, Coulomb-mediated electron hole dynamics more important for rutile than for anatase because formation of the
need to be taken into account in their design [128]. superoxide radical anion is more difficult on Pt-free rutile than on
To improve the photocatalytic activity, the close electronic Pt-free anatase (the higher energy (by ∼0.2 eV) of the anatase con-
interaction between the noble metal catalyst and the photocatalyst duction band may facilitate a more rapid transfer of electrons to O2 ).
is very important. Bahnemann and coworkers studied two kinds Emilio et al. have investigated three commercial TiO2 photocatalyst
of Pt-decorated TiO2 photocatalysts [130]. One has been prepared powders (Degussa P25, Sachtleben UV100, and Millenium PC50)
by photodeposition of Pt yielding Pt clusters on the TiO2 surface, and their platinized forms by the time-resolved microwave con-
while the other one has been prepared by the mixing of colloidal Pt ductivity (TRMC) method to follow their charge-carrier dynamics
with excess colloidal TiO2 , i.e., the Pt particles should rather be sur- and to relate it to the photocatalytic activity [133]. They found that
rounded by TiO2 particles. The in situ growth of Pt clusters during platinization has a distinct influence on these commercial products,
photodeposition leaves little doubt that the electronic interaction decreasing globally the activity of P25 and increasing the activity
between Pt and TiO2 in photodeposited Pt islands is stronger than in of PC50 and UV100. The platinization also had a distinct influ-
physically mixed ones. It is not surprising that photochemical pla- ence on the charge-carrier lifetimes. The authors proposed that in
tinization of the TiO2 particles yields a more efficient photocatalyst the long-time range, platinization improves the separation of the
for the photocatalytic degradation of methanol than the physical charge carriers in PC50, has almost no effect on P25, and acceler-
mixing of the colloidal components of Pt and TiO2 . ates the decay of electrons in UV100. The activity of P25 seems to
The catalytic activities of noble metal deposited TiO2 systems be determined by short-time range effects, while, in UV100 and
depend on many aspects, such as the particle size of the noble metal, mostly in PC50, the effects in the long-time range are the dominant
the coating amount, the type of noble metal, the crystalline phase ones [133].
of TiO2 , etc. Goodman and coworkers have investigated the unusual
size dependence of the low-temperature catalytic oxidation of car- 5.2. Plasmonic structures
bon monoxide by preparing gold clusters ranging in diameter from
1 to 6 nanometers on single crystalline surfaces of titania [131]. Plasmonic metallic nanostructures exhibit strong interaction
They proposed that the structure sensitivity of this reaction on gold with resonant photons through an excitation of the surface plas-
clusters supported on titania is related to a quantum size effect with mon resonance (SPR). Recently, growing attention has been paid
respect to the thickness of the gold islands, with islands that were on employing plasmonic structures to improve the efficiencies of
composed of two layers of gold being most effective for catalyzing photocatalysis, solar-to-H2 conversion, and solar cells [134–143].
the oxidation of carbon monoxide. Kamat and coworkers probed It is well known that a crucial obstacle limiting the efficiency of
the transfer of electrons to Au nanoparticles with different parti- photocatalysts is the high rate of charge-carrier recombination,
cle size by exciting TiO2 nanoparticles under steady-state and laser which is attributed to the large penetration depth of photons as
pulse excitation, respectively [126]. Equilibration with the C60 /C60 − compared with the short diffusion distance of the photogenerated
redox couple was used to determine the apparent Fermi level of charge carriers. One method to reduce the recombination rate is to
the TiO2 –Au composite system. The shift in the apparent Fermi minimize the distance between the charge carrier generating loca-
level of the TiO2 –Au composite was found to be size-dependent: tion and the photocatalyst surface. Theoretically, recombination
20 mV for 8-nm diameter, 40 mV for 5-nm and 60 mV for 3-nm can be reduced, if the light absorption is confined to the near-
gold nanoparticles. surface region of a semiconductor so that the charge carrier travel
The enhancement effect of noble metal deposition is also distance is shortened. Plasmonic metal nanostructures with sur-
found to be different on TiO2 with different crystalline phase. face plasmon resonance can act as an antenna that localizes optical
Egerton et al. studied the effect of deposition of Pt on the UV energy and changes the charge carrier generating location. The
L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276 273

Fig. 11. TEM view. Cross section of TiO2 film on Ag/SiO2 core–shell on a SiO2 sub-
strate.
Ref. [134] Copyright 2008 ACS.

interaction of localized electric fields around the plasmonic metal


particles with a neighboring semiconductor enables the selective
formation of electron/hole pairs in the near-surface region of the
semiconductor. This can result in a decrease in the charge-carrier
Fig. 12. Average electric field enhancement around a Ag cube with an edge length
recombination rate and in an increase in the overall efficiency of of 120 nm as a function of the distance d from the cube, as calculated using FDTD
the photocatalytic reaction. simulations. Inset: local enhancement of the electric field calculated from an FDTD
In 2008, Awazu et al. investigated a plasmonic photocatalyst simulation of a 120 nm Ag cube in water.
consisting of silver nanoparticles embedded in titanium dioxide Ref. [144] Copyright 2011 ACS.
[134]. They proposed that the photocatalytic behavior of TiO2
would be greatly boosted if it is assisted by the enhanced near-
photocatalytic efficiency. These effects will, however, not be dis-
field amplitudes of localized surface plasmons (LSP) of the silver
cussed in the current work. Recently, a few review articles have
nanoparticles. They found that the key to enable plasmonic pho-
been published on the topic of plasmonic structures for efficient
tocatalysis is to deposit TiO2 on a NP-structure comprising an Ag
solar energy conversion [137,145].
core covered with a silica (SiO2 ) shell to prevent oxidation of Ag
by direct contact with TiO2 (as shown in Fig. 11). By Mie scattering
theory they estimated the most appropriate diameter for Ag NPs 5.3. Graphenic carbon coatings
and thickness for the SiO2 shell giving rise to LSP in the near UV.
The measured photocatalytic activity under near UV illumination of Many experimental studies have been devoted to reduce the
such a plasmonic photocatalyst, monitored by the decomposition recombination of charge carriers by coupling the photocatalysts
of methylene blue, was enhanced by a factor of 7. The enhancement with suitable materials, such as noble metals, semiconductors
of the photocatalytic activity increases with a decreased thickness [146,147], etc. Recently, it has been reported that combining con-
of the SiO2 shell. Photocatalytic activity with 100 nm thick SiO2 was jugative ␲ structure materials with the semiconductor particles can
observed to be close to that of pure TiO2 . This fact implies that the form efficient photocatalysts, such as C60 [148–150], carbon nano-
increase in photocatalytic activity was obtained exclusively by the tubes [151–155], and graphene [156–170]. The hybridization of a
LSP resonance from the Ag nanoparticles [134]. photocatalyst with conjugative ␲ structure materials possessing
Linic and coworkers studied the photocatalytic water splitting good electrical conductivity could reduce the recombination rate
on composite plasmonic-metal/semiconductor photoelectrodes of the charge carriers, and hence increase the photocatalytic effi-
[144]. They found that plasmonic Ag nanostructures can be ciency. Among these conjugative ␲ structure materials, graphene
employed as building blocks to create composite plasmonic- has attracted the most attention due to its unique properties such
metal/semiconductor photoelectrocatalysts that exhibit enhanced as high conductivity and electron mobility. The incorporation of
water splitting performance relative to the semiconductor alone. graphene into photocatalysts can additionally equip them with the
This enhancement was attributed to increased local electric fields unique properties of graphene and possibly induce new properties,
in the neighborhood of the Ag nanostructures due to the formation such as a high dye adsorption capacity, an extended light absorp-
of resonant surface plasmon (SP) states. Fig. 12 shows the FDTD- tion range, and an enhanced charge separation and transportation
calculated field enhancements from a 120 nm Ag cube in water. properties, thus enhancing the overall photocatalytic performance
The electric fields are spatially nonhomogenous, with the highest [156].
field strength obtained in the proximity of the nanostructures. It Employing a hydrothermal process, Zhu and coworkers found
is suggested that SP-induced e− /h+ pair formation should be most that surface hybridization of TiO2 with graphene-like carbon layers
pronounced in the part of the semiconductor that is the closest to of a few molecular layers thickness yields efficient photocat-
Ag, i.e., near the surface of the semiconductor particles (essentially alysts [169]. Photoelectrochemical measurements confirmed an
at the semiconductor–liquid interface). The authors proposed two electronic interaction between TiO2 and the graphite-like carbon.
advantages of forming e− /h+ pairs near the semiconductor surface They observed that a TiO2 photocatalyst with a carbon shell of
rather than in the bulk: (i) the charge carriers are readily separated three molecular layers thickness (∼1 nm) shows the highest pho-
from each other under the influence of the surface potential and tocatalytic activity which is about two times higher than that of
(ii) the charge carriers have a shorter distance to migrate in order Degussa P25 TiO2 under UV-light irradiation. The enhanced pho-
to reach the surface. This results in a decrease in the charge–carrier tocatalytic activity under UV irradiation is attributed to the high
recombination rate and increases the water splitting rate on com- migration efficiency of the photoinduced electrons at the graphene-
posite photocatalysts that are composed of a plasmonic metal and like carbon/TiO2 interface, as shown in Fig. 13. In addition, a high
a semiconductor [144]. photocatalytic activity under visible light irradiation is observed
Besides the surface plasmon effect in reducing the charge car- after graphene-like carbon hybridization [169].
rier recombination by selective formation of e− /h+ pairs in the Graphene hybridized ZnO photocatalyst particles also showed
semiconductor, other effects of plasmonic metallic structures such enhanced photocatalytic activity for the degradation of organic
as, scattering, anti-reflection, hot electron/hot hole transfer from dyes [170]. It was found that the degree of photocatalytic activ-
metal to semiconductor are also of great interest in enhancing the ity enhancement strongly depended on the coverage of graphene
274 L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276

enhancement is still not clear. There are only few studies concern-
ing the charge transfer kinetics in the graphene/TiO2 systems. More
work is necessary to clarify the charge transfer mechanism between
graphene and TiO2 .

6. Conclusions

Charge carrier transfer processes are very important and play


a vital role in a photocatalysis reaction. The fundamental study of
the dynamics of the charge transfer process is thus crucial from the
viewpoint of developing efficient photocatalysis systems for large-
scale industrialization. The current work mainly reviews recent
efforts on understanding the charge transfer kinetics in the photo-
Fig. 13. TiO2 /graphene layers composite as efficient photocatalyst. catalytic process. Some fundamental aspects involved in the charge
Ref. [169] Copyright 2008 Wiley. transfer process, such as, charge carrier generation, charge car-
rier trapping, charge carrier recombination, and electron and hole
transfer are discussed based on the results published in the past
decades. Some recent studies focusing on the enhancement of the
photocatalytic efficiency by improving the charge carrier transfer
and separation are also reviewed in this work. Noble metal load-
ing, plasmonic structure, and graphene loading have been found to
be efficient methods to improve charge carrier separation and to
suppress charge carrier recombination. Although there have been
significant advances in the research of charge transfer dynamics,
there are still many processes not fully understood, especially on
the molecular-level. There are, for example, hardly any studies
associated with electron and hole transfer kinetics in photocatalytic
reactions on single crystal TiO2 surfaces. Most researchers have
studied the charge transfer kinetics on a very short timescale, while
the charge transfer on a more extended timescale is still unclear.
This review has highlighted the importance of charge transfer pro-
cesses in photocatalytic reactions the understanding of which can
provide possibilities to significantly improve photocatalytic effi-
Fig. 14. Fluorescence decay traces of 1 mM ZnO with varying GO concentrations. ciencies.
Inset shows the dependence of lifetime on GO concentration.
Ref. [171] Copyright 2009 ACS.
Acknowledgements

on the surface of the employed ZnO nanoparticles. The sample of


L.Z. acknowledges the Alexander von Humboldt (AvH) Founda-
2 wt% graphene hybridized ZnO showed the highest photocatalytic
tion for granting him a research fellowship. Financial support from
activity, which was about 4 times higher than that of pristine ZnO.
the Deutsche Forschungsgemeinschaft (DFG) is gratefully acknowl-
The enhancement of the photocatalytic activity was attributed to
edged (Grant no. BA 1137/5-1).
an inhibited charge carrier recombination due to the electronic
interaction between ZnO and graphene. At higher loading with
graphene, the photocatalytic activity is decreased due to a light References
shielding effect of graphene on the surface [170].
[1] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Chem. Rev. 95 (1995)
To study the electron transfer from the semiconductor to 69–96.
graphene, Kamat and coworkers have investigated the emission [2] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293 (2001) 269–271.
decay of a suspension of ZnO nanoparticles at varying concentra- [3] A.L. Linsebigler, G. Lu, J.T. Yates Jr., Chem. Rev. 95 (1995) 735–758.
[4] A. Fujishima, T.N. Rao, D.A. Tryk, J. Photochem. Photobiol. C 1 (2000) 1–21.
tions of graphene oxide (GO). As shown in Fig. 14 the resulting [5] A. Fujishima, X. Zhang, D.A. Tryk, Surf. Sci. Rep. 63 (2008) 515–582.
fluorescence spectra were fitted to derive the respective decay [6] O.M. Alfano, D. Bahnemann, A.E. Cassano, R. Dillert, R. Goslich, Catal. Today
time constants, which can be used to calculate average lifetimes 58 (2000) 199–230.
[7] M. Ni, M.K.H. Leung, D.Y.C. Leung, K. Sumathy, Renew. Sustain. Energy Rev.
of the fluorescent, i.e., excited state. The inset of Fig. 14 shows 11 (2007) 401–425.
the decrease in average lifetime as a function of the GO concen- [8] U.I. Gaya, A.H. Abdullah, J. Photochem. Photobiol. C 9 (2008) 1–12.
tration. A decrease in the average lifetime of the ZnO emission [9] R.W. Matthews, J. Catal. 111 (1988) 264–272.
[10] R.W. Matthews, J. Phys. Chem. 91 (1987) 3328–3333.
from 30 to 14 ns was observed with increasing concentration of [11] M.A. Henderson, Surf. Sci. Rep. 66 (2011) 185–297.
GO. The short component shows a nearly 20-fold decrease in life- [12] T. Tachikawa, M. Fujitsuka, T. Majima, J. Phys. Chem. C 111 (2007) 5259–5275.
time (2.98–0.67 ns) when the GO concentration is increased to [13] C. Kormann, D.W. Bahnemann, M.R. Hoffmann, J. Phys. Chem. 92 (1988)
5196–5201.
0.25 mg/mL [171]. As a result of the efficient electron-transport
[14] L.E. Brus, J. Chem. Phys. 80 (1984) 4403.
properties, graphene also significantly quenched the fluorescence [15] R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki, Y. Taga, Science 293 (2001) 269.
of CdS quantum dots (QDs) decorated on it. Cao et al. employed [16] T. Umebayashi, T. Yamaki, H. Itoh, K. Asai, Appl. Phys. Lett. 81 (2002) 454.
[17] T. Umebayashi, T. Yamaki, S. Tanaka, K. Asai, Chem. Lett. 32 (2003) 330–331.
time-resolved fluorescence spectroscopy to monitor the emission
[18] Y. Nakano, T. Morikawa, T. Ohwaki, Y. Taga, Appl. Phys. Lett. 87 (2005) 052111.
lifetimes of free CdS QDs and G-CdS [172]. They observed an ultra- [19] H. Irie, Y. Watanabe, K. Hashimoto, Chem. Lett. 32 (2003) 772–773.
fast process (5 ps) of the electron transfer from the excited CdS to [20] S. Sakthivel, H. Kisch, Angew. Chem. Int. Ed. 42 (2003) 4908–4911.
the graphene matrices. [21] S.U.M. Khan, M. Al-Shahry, W.B. Ingler, Science 297 (2002) 2243.
[22] J.O. Carneiro, V. Teixeira, A. Portinha, A. Magalhaes, P. Coutinho, C.J. Tavares,
Although considerable progress on graphene modified pho- R. Newton, Mater. Sci. Eng. B 138 (2007) 144–150.
tocatalysis has been achieved, the detailed mechanism of the [23] J. Yu, M. Zhou, H. Yu, Q. Zhang, Y. Yu, Mater. Chem. Phys. 95 (2006) 193–196.
L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276 275

[24] X.Z. Li, F.B. Li, C.L. Yang, W.K. Ge, J. Photochem. Photobiol. A 141 (2001) [75] Y. Wang, K. Hang, N.A. Anderson, T. Lian, J. Phys. Chem. B 107 (2003)
209–217. 9434–9440.
[25] Y.C. Lee, Y.P. Hong, H.Y. Lee, H. Kim, Y.J. Jung, K.H. Ko, H.S. Jung, K.S. Hong, J. [76] R.L. Wells, W.L. Gladfelter, J. Clust. Sci. 8 (1997) 217–238.
Colloid Interface Sci. 267 (2003) 127–131. [77] J.M. Herrmann, J. Disdier, P. Pichat, J. Phys. Chem. 90 (1986) 6028–6034.
[26] G.M. Turner, M.C. Beard, C.A. Schmuttenmaer, J. Phys. Chem. B 106 (2002) [78] G.A. Somorjai, Y.G. Borodko, Catal. Lett. 76 (2001) 1–5.
11716–11719. [79] Z. Ma, F. Zaera, in: R.B. King (Ed.), Encyclopedia of Inorganic Chemistry, John
[27] T. Morishita, A. Hibara, T. Sawada, I. Tsuyumoto, J. Phys. Chem. B 103 (1999) Wiley & Sons, New York, 2005.
5984–5987. [80] M. Che, C.O. Bennett, Adv. Catal. 36 (1989) 55–172.
[28] M.A. Grela, M.A. Brusa, A.J. Colussi, J. Phys. Chem. B 101 (1997) 10986–10989. [81] G.C. Bond, Acc. Chem. Res. 26 (1993) 490–495.
[29] M.A. Grela, A.J. Colussi, J. Phys. Chem. B 103 (1999) 2614–2619. [82] G.A. Somorjai, Catal. Lett. 7 (1990) 169–182.
[30] A. Emeline, A. Salinaro, N. Serpone, J. Phys. Chem. B 104 (2000) 11202–11210. [83] J.F. Lambert, M. Che, J. Mol. Catal. A: Chem. 162 (2000) 5–18.
[31] Y. Tamaki, A. Furube, M. Murai, K. Hara, R. Katoh, M. Tachiya, J. Am. Chem. [84] H. Reiche, W.W. Dunn, A.J. Bard, J. Phys. Chem. 83 (1979) 2248–2251.
Soc. 128 (2006) 416–417. [85] H.H. Mohamed, C.B. Mendive, R. Dillert, D.W. Bahnemann, J. Phys. Chem. A
[32] D. Bahnemann, A. Henglein, J. Lilie, L. Spanhel, J. Phys. Chem. 88 (1984) 115 (2011) 2139–2147.
709–711. [86] H.H. Mohamed, R. Dillert, D.W. Bahnemann, J. Phys. Chem. C 115 (2011)
[33] D. Bahnemann, A. Henglein, L. Spanhel, Faraday Discuss. Chem. Soc. 78 (1984) 12163–12172.
151–163. [87] H.H. Mohamed, R. Dillert, D.W. Bahnemann, Chem. Eur. J. 18 (2012)
[34] S. Leytner, J.T. Hupp, Chem. Phys. Lett. 330 (2000) 231–236. 4314–4321.
[35] Y. Tamaki, A. Furube, M. Murai, K. Hara, R. Katoh, M. Tachiya, Phys. Chem. [88] E. Vrachnou, M. Grätzel, A.J. McEvoy, J. Electroanal. Chem. 258 (1989)
Chem. Phys. 9 (2007) 1453–1460. 193–205.
[36] A. Furube, T. Asahi, H. Masuhara, H. Yamashita, M. Anpo, J. Phys. Chem. B 103 [89] S.O. Obare, T. Ito, M.H. Balfour, G.J. Meyer, Nano Lett. 3 (2003) 1151–1153.
(1999) 3120–3127. [90] S.O. Obare, T. Ito, G.J. Meyer, J. Am. Chem. Soc. 128 (2006) 712–713.
[37] D.P. Colombo Jr., R.M. Bowman, J. Phys. Chem. 100 (1996) 18445–18449. [91] T. Ito, G.J. Meyer, Environ. Eng. Sci. 24 (2007) 31–44.
[38] D.P. Colombo Jr., R.M. Bowman, J. Phys. Chem. 99 (1995) 11752–11756. [92] R. Gao, A. Safrany, J. Rabani, Radiat. Phys. Chem. 67 (2003) 25–39.
[39] N.J. Cherepy, G.P. Smestad, M. Grätzel, J.Z. Zhang, J. Phys. Chem. B 101 (1997) [93] K.T. Ranjit, B. Viswanathan, J. Photochem. Photobiol. A 108 (1997) 73–78.
9342–9351. [94] K.T. Ranjit, B. Viswanathan, T.K. Varadarajan, J. Mater. Sci. Lett. 15 (1996)
[40] G. Rothenberger, J. Moser, M. Grätzel, N. Serpone, D.K. Sharma, J. Am. Chem. 874–877.
Soc. 107 (1985) 8054–8059. [95] F. Zhang, Y. Pi, J. Cui, Y. Yang, X. Zhang, N. Guan, J. Phys. Chem. C 111 (2007)
[41] N. Serpone, D. Lawless, R. Khairutdinov, E. Pelizzetti, J. Phys. Chem. 99 (1995) 3756–3761.
16655–16661. [96] M. Halmann, K. Zuckerman, J. Chem. Soc. Chem. Commun. (1986) 455–457.
[42] X. Yang, N. Tamai, Phys. Chem. Chem. Phys. 3 (2001) 3393–3398. [97] Z. Kasarevic-Popovic, D. Behar, J. Rabani, J. Phys. Chem. B 108 (2004)
[43] S. Ikeda, N. Sugiyama, S. Murakami, H. Kominami, Y. Kera, H. Noguchi, K. 20291–20295.
Uosaki, T. Torimoto, B. Ohtani, Phys. Chem. Chem. Phys. 5 (2003) 778–783. [98] F.H. Hussein, R. Rudham, J. Chem. Soc. Faraday Trans. 1 (83) (1987) 1631–1639.
[44] A.I. Kuznetsov, O. Kameneva, A. Alexandrov, N. Bityurin, P. Marteau, K. Chhor, [99] D. Behar, J. Rabani, J. Phys. Chem. B 110 (2006) 8750–8755.
C. Sanchez, A. Kanaev, Phys. Rev. E 71 (2005) 021403. [100] P. Robertson, D. Bahnemann, J. Robertson, F. Wood, Photocatalytic Detoxifi-
[45] Q. Shen, K. Katayama, T. Sawada, M. Yamaguchi, Y. Kumagai, T. Toyoda, Chem. cation of Water and Air, Springer-Verlag, Berlin Heidelberg, 2005.
Phys. Lett. 419 (2006) 464–468. [101] M. Simic, P. Neta, E. Hayon, J. Phys. Chem. 73 (1969) 3794–3800.
[46] G. Boschloo, D. Fitzmaurice, J. Phys. Chem. B 103 (1999) 2228–2231. [102] M. Anpo, T. Shima, Y. Kubokawa, Chem. Lett. (1985) 1799–1802.
[47] R.B. Lauer, R.R. Addiss, A.K. Ghosh, J. Appl. Phys. 42 (1971) 3508–3512. [103] T.D. Bui, A. Kimura, S. Ikeda, M. Matsumura, J. Am. Chem. Soc. 132 (2010)
[48] G. Redmond, D. Fitzmaurice, M. Grätzel, J. Phys. Chem. 97 (1993) 6951–6954. 8453–8458.
[49] N. Serpone, D. Lawless, R. Khairutdinov, J. Phys. Chem. 99 (1995) [104] A. Imanishi, T. Okamura, N. Ohashi, R. Nakamura, Y. Nakato, J. Am. Chem. Soc.
16646–16654. 129 (2007) 11569–11578.
[50] D. Lawless, N. Serpone, D. Meisel, J. Phys. Chem. 95 (1991) 5166–5170. [105] J.A. Anderson, Catal. Today 175 (2011) 316–321.
[51] M.A. Henderson, J.M. White, H. Uetsuka, H. Onishi, J. Am. Chem. Soc. 125 [106] I. Bannat, K. Wessels, T. Oekermann, J. Rathousky, D. Bahnemann, M. Wark,
(2003) 14974–14975. Chem. Mater. 21 (2009) 1645–1653.
[52] D.W. Bahnemann, M. Hilgendorff, R. Memming, J. Phys. Chem. B 101 (1997) [107] B. Cojocaru, S. Neatu, E. Sacaliuc-Parvulescu, F. Levy, V.I. Parvulescu, H. Garcia,
4265–4275. Appl. Catal. B 107 (2011) 140–149.
[53] S.H. Szczepankiewicz, J.A. Moss, M.R. Hoffmann, J. Phys. Chem. B 106 (2002) [108] D.M. Fouad, M.B. Mohamed, Nanotechnology 22 (2011).
2922–2927. [109] A.A. Ismail, D.W. Bahnemann, I. Bannat, M. Wark, J. Phys. Chem. C 113 (2009)
[54] A.M. Peiro, C. Colombo, G. Doyle, J. Nelson, A. Mills, J.R. Durrant, J. Phys. Chem. 7429–7435.
B 110 (2006) 23255–23263. [110] O. Rosseler, M.V. Shankar, M.K.L. Du, L. Schmidlin, N. Keller, V. Keller, J. Catal.
[55] C.P. Kumar, N.O. Gopal, T.C. Wang, M.S. Wong, S.C. Ke, J. Phys. Chem. B 110 269 (2010) 179–190.
(2006) 5223–5229. [111] C.G. Silva, R. Juarez, T. Marino, R. Molinari, H. Garcia, J. Am. Chem. Soc. 133
[56] S.H. Szczepankiewicz, J.A. Moss, M.R. Hoffmann, J. Phys. Chem. B 106 (2002) (2011) 595–602.
7654–7658. [112] Q. Xiang, G.F. Meng, H.B. Zhao, Y. Zhang, H. Li, W.J. Ma, J.Q. Xu, J. Phys. Chem.
[57] S.T. Martin, H. Herrmann, W. Choi, M.R. Hoffmann, J. Chem. Soc. Faraday Trans. C 114 (2010) 2049–2055.
90 (1994) 3315–3322. [113] S. Hwang, M.C. Lee, W. Choi, Appl. Catal. B 46 (2003) 49–63.
[58] S.H. Szczepankiewicz, A.J. Colussi, M.R. Hoffmann, J. Phys. Chem. B 104 (2000) [114] K.L. Miller, C.W. Lee, J.L. Falconer, J.W. Medlin, J. Catal. 275 (2010) 294–299.
9842–9850. [115] H.Q. Sun, R. Ullah, S.H. Chong, H.M. Ang, M.O. Tade, S.B. Wang, Appl. Catal. B
[59] C.D. Jaeger, A.J. Bard, J. Phys. Chem. 83 (1979) 3146–3152. 108 (2011) 127–133.
[60] R.F. Howe, M. Grätzel, J. Phys. Chem. 89 (1985) 4495–4499. [116] K. Koci, K. Mateju, L. Obalova, S. Krejcikova, Z. Lacny, D. Placha, L. Capek, A.
[61] T. Yoshihara, Y. Tamaki, A. Furube, M. Murai, K. Hara, R. Katoh, Chem. Phys. Hospodkova, O. Solcova, Appl. Catal. B 96 (2010) 239–244.
Lett. 438 (2007) 268–273. [117] A.V. Korzhak, N.I. Ermokhina, A.L. Stroyuk, V.K. Bukhtiyarov, A.E. Raevskaya,
[62] T. Berger, M. Sterrer, O. Diwald, E. Kn zinger, D. Panayotov, T.L. Thompson, J.T. V.I. Litvin, S.Y. Kuchmiy, V.G. Ilyin, P.A. Manorik, J. Photochem. Photobiol. A –
Yates Jr., J. Phys. Chem. B 109 (2005) 6061–6068. Chem. 198 (2008) 126–134.
[63] S. Kerisit, N.A. Deskins, K.M. Rosso, M. Dupuis, J. Phys. Chem. C 112 (2008) [118] A. Kubacka, M. Ferrer, A. Martinez-Arias, M. Fernandez-Garcia, Appl. Catal. B
7678–7688. 84 (2008) 87–93.
[64] O.I. Micic, Y. Zhang, K.R. Cromack, A.D. Trifunac, M.C. Thurnauer, J. Phys. Chem. [119] Y.C. Lin, H.S. Lee, J. Hazard. Mater. 179 (2010) 462–470.
97 (1993) 7277–7283. [120] A.V. Rupa, D. Manikandan, D. Divakar, T. Sivakumar, J. Hazard. Mater. 147
[65] D.C. Hurum, A.G. Agrios, K.A. Gray, T. Rajh, M.C. Thurnauer, J. Phys. Chem. B (2007) 906–913.
107 (2003) 4545–4549. [121] E. Szabo-Bardos, E. Petervari, V. El-Zein, A. Horvath, J. Photochem. Photobiol.
[66] H. Tang, K. Prasad, R. Sanjines, P.E. Schmid, F. Levy, J. Appl. Phys. 75 (1994) A – Chem. 184 (2006) 221–227.
2042–2047. [122] R. van Grieken, J. Marugan, C. Sordo, P. Martinez, C. Pablos, Appl. Catal. B 93
[67] H. Tang, F. Levy, H. Berger, P.E. Schmid, Phys. Rev. B 52 (1995) 7771. (2009) 112–118.
[68] D. Philip Colombo, K.A. Roussel, J. Saeh, D.E. Skinner, J.J. Cavaleri, R.M. Bow- [123] B.F. Xin, Z.Y. Ren, H.Y. Hu, X.Y. Zhang, C.L. Dong, K.Y. Shi, L.Q. Jing, H.G. Fu,
man, Chem. Phys. Lett. 232 (1995) 207–214. Appl. Surf. Sci. 252 (2005) 2050–2055.
[69] M.A. Grela, A.J. Colussi, J. Phys. Chem. 100 (1996) 18214–18221. [124] B. Kraeutler, A.J. Bard, J. Am. Chem. Soc. 100 (1978) 4317–4318.
[70] A. Yamakata, T.a. Ishibashi, H. Onishi, Chem. Phys. Lett. 333 (2001) [125] Y. Nosaka, Y. Ishizuka, H. Miyama, Ber. Buns. Phys. Chem. 90 (1986)
271–277. 1199–1204.
[71] T. Yoshihara, R. Katoh, A. Furube, Y. Tamaki, M. Murai, K. Hara, S. Murata, H. [126] V. Subramanian, E.E. Wolf, P.V. Kamat, J. Am. Chem. Soc. 126 (2004)
Arakawa, M. Tachiya, J. Phys. Chem. B 108 (2004) 3817–3823. 4943–4950.
[72] R. Katoh, A. Furube, J. Phys. Chem. Lett. 2 (2011) 1888–1891. [127] S. Sarkar, A. Makhal, T. Bora, S. Baruah, J. Dutta, S.K. Pal, Phys. Chem. Chem.
[73] A. Yamakata, T. Ishibashi, H. Onishi, J. Phys. Chem. B 105 (2001) 7258–7262. Phys. 13 (2011) 12488.
[74] A.V. Emeline, V.K. Ryabchuk, N. Serpone, J. Phys. Chem. B 109 (2005) [128] M.J. Berr, A. Vaneski, C. Mauser, S. Fischbach, A.S. Susha, A.L. Rogach, F. Jackel,
18515–18521. J. Feldmann, Small 8 (2012) 291.
276 L. Zhang et al. / Journal of Photochemistry and Photobiology C: Photochemistry Reviews 13 (2012) 263–276

[129] K. Iwata, T. Takaya, H. Hamaguchi, A. Yamakata, T. Ishibashi, H. Onishi, H. [150] L.W. Zhang, Y.J. Wang, T.G. Xu, S.B. Zhu, Y.F. Zhu, J. Mol. Catal. A: Chem. 331
Kuroda, J. Phys. Chem. B 108 (2004) 20233–20239. (2010) 7–14.
[130] C. Wang, R. Pagel, D.W. Bahnemann, J.K. Dohrmann, J. Phys. Chem. B 108 [151] K. Woan, G. Pyrgiotakis, W. Sigmund, Adv. Mater. 21 (2009) 2233–2239.
(2004) 14082–14092. [152] M. Daranyi, T. Csesznok, A. Kukovecz, Z. Konya, I. Kiricsi, P.M. Ajayan, R. Vajtai,
[131] M. Valden, X. Lai, D.W. Goodman, Science 281 (1998) 1647–1650. Nanotechnology 22 (2011).
[132] T.A. Egerton, H. Purnama, J.A. Mattinson, J. Photochem. Photobiol. A. Chem. [153] Z.R. Tang, F. Li, Y.H. Zhang, X.Z. Fu, Y.J. Xu, J. Phys. Chem. C 115 (2011)
224 (2011) 31–37. 7880–7886.
[133] C.A. Emilio, I. Marta, M. Kunst, M. Bouchard, C. Colbeau-Justin, Langmuir 22 [154] C.G. Silva, J.L. Faria, Chemsuschem 3 (2010) 609–618.
(2006) 3606–3613. [155] Y. Ou, J.D. Lin, S.M. Fang, D.W. Liao, Chem. Phys. Lett. 429 (2006) 199–203.
[134] K. Awazu, M. Fujimaki, C. Rockstuhl, J. Tominaga, H. Murakami, Y. Ohki, N. [156] Q.J. Xiang, J.G. Yu, M. Jaroniec, Chem. Soc. Rev. 41 (2012) 782–796.
Yoshida, T. Watanabe, J. Am. Chem. Soc. 130 (2008) 1676–1680. [157] Z. Hu, Y.D. Huang, S.F. Sun, W.C. Guan, Y.H. Yao, P.Y. Tang, C.Y. Li, Carbon 50
[135] W.B. Hou, Z.W. Liu, P. Pavaskar, W.H. Hung, S.B. Cronin, J. Catal. 277 (2011) (2012) 994–1004.
149–153. [158] P.D. Tran, L.H. Wong, J. Barber, J.S.C. Loo, Energy Environ. Sci. 5 (2012)
[136] D.B. Ingram, P. Christopher, J.L. Bauer, S. Linic, ACS Catal. 1 (2011) 1441–1447. 5902–5918.
[137] S. Linic, P. Christopher, D.B. Ingram, Nat. Mater 10 (2011) 911–921. [159] J.S. Lee, K.H. You, C.B. Park, Adv. Mater. 24 (2012) 1084–1088.
[138] Z.W. Liu, W.B. Hou, P. Pavaskar, M. Aykol, S.B. Cronin, Nano Lett. 11 (2011) [160] S. Ghasemi, S.R. Setayesh, A. Habibi-Yangjeh, M.R. Hormozi-Nezhad, M.R.
1111–1116. Gholami, J. Hazard. Mater. 199 (2012) 170–178.
[139] Y. Lu, H.T. Yu, S. Chen, X. Quan, H.M. Zhao, Environ. Sci. Technol. 46 (2012) [161] Y.S. Fu, H.Q. Chen, X.Q. Sun, X. Wang, Appl. Catal. B 111 (2012) 280–287.
1724–1730. [162] H.I. Kim, G.H. Moon, D. Monllor-Satoca, Y. Park, W. Choi, J. Phys. Chem. C 116
[140] S. Mubeen, G. Hernandez-Sosa, D. Moses, J. Lee, M. Moskovits, Nano Lett. 11 (2012) 1535–1543.
(2011) 5548–5552. [163] D.L. Zhao, G.D. Sheng, C.L. Chen, X.K. Wang, Appl. Catal. B 111 (2012) 303–308.
[141] S.M. Sun, W.Z. Wang, L. Zhang, M. Shang, L. Wang, Catal. Commun. 11 (2009) [164] N. Li, G. Liu, C. Zhen, F. Li, L.L. Zhang, H.M. Cheng, Adv. Funct. Mater. 21 (2011)
290–293. 1717–1722.
[142] J.G. Yu, G.P. Dai, B.B. Huang, J. Phys. Chem. C 113 (2009) 16394–16401. [165] J. Du, X.Y. Lai, N.L. Yang, J. Zhai, D. Kisailus, F.B. Su, D. Wang, L. Jiang, ACS Nano
[143] J.B. Zhou, Y. Cheng, J.G. Yu, J. Photochem. Photobiol. A 223 (2011) 82–87. 5 (2011) 590–596.
[144] D.B. Ingram, S. Linic, J. Am. Chem. Soc. 133 (2011) 5202–5205. [166] Y.H. Zhang, Z.R. Tang, X.Z. Fu, Y.J. Xu, ACS Nano 4 (2010) 7303–7314.
[145] S.C. Warren, E. Thimsen, Energy Environ. Sci. 5 (2012) 5133–5146. [167] H. Zhang, X.J. Lv, Y.M. Li, Y. Wang, J.H. Li, ACS Nano 4 (2010) 380–386.
[146] S.H. Elder, F.M. Cot, Y. Su, S.M. Heald, A.M. Tyryshkin, M.K. Bowman, Y. Gao, [168] G. Williams, B. Seger, P.V. Kamat, ACS Nano 2 (2008) 1487–1491.
A.G. Joly, M.L. Balmer, A.C. Kolwaite, K.A. Magrini, D.M. Blake, J. Am. Chem. [169] L.W. Zhang, H.B. Fu, Y.F. Zhu, Adv. Funct. Mater. 18 (2008) 2180–2189.
Soc. 122 (2000) 5138–5146. [170] T.G. Xu, L.W. Zhang, H.Y. Cheng, Y.F. Zhu, Appl. Catal. B 101 (2011)
[147] T. Tatsuma, S. Saitoh, P. Ngaotrakanwiwat, Y. Ohko, A. Fujishima, Langmuir 382–387.
18 (2002) 7777–7779. [171] G. Williams, P.V. Kamat, Langmuir 25 (2009) 13869–13873.
[148] H.B. Fu, T.G. Xu, S.B. Zhu, Y.F. Zhu, Environ. Sci. Technol. 42 (2008) [172] A.N. Cao, Z. Liu, S.S. Chu, M.H. Wu, Z.M. Ye, Z.W. Cai, Y.L. Chang, S.F. Wang,
8064–8069. Q.H. Gong, Y.F. Liu, Adv. Mater. 22 (2010) 103.
[149] S.B. Zhu, T.G. Xu, H.B. Fu, J.C. Zhao, Y.F. Zhu, Environ. Sci. Technol. 41 (2007)
6234–6239.

You might also like