(2011) Crystal Facet Engineering of Semiconductor Photocatalysts (Motivations, Advances and Unique Properties) - Gang Liu, Et Al

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

View Article Online / Journal Homepage / Table of Contents for this issue

ChemComm Dynamic Article Links

Cite this: Chem. Commun., 2011, 47, 6763–6783

www.rsc.org/chemcomm FEATURE ARTICLE


Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

Crystal facet engineering of semiconductor photocatalysts: motivations,


advances and unique properties
Gang Liu,a Jimmy C. Yu,*b Gao Qing (Max) Luc and Hui-Ming Cheng*a
Received 2nd February 2011, Accepted 8th March 2011
DOI: 10.1039/c1cc10665a

Crystal facet engineering of semiconductors has become an important strategy for fine-tuning the
physicochemical properties and thus optimizing the reactivity and selectivity of photocatalysts.
In this review, we present the basic strategies for crystal facet engineering of photocatalysts and
describe the recent advances in synthesizing faceted photocatalysts, in particular TiO2 crystals.
The unique properties of faceted photocatalysts are discussed in relation to anisotropic corrosion,
interaction dependence of adsorbates, photocatalytic selectivity, photo-reduction and oxidation
sites, and photocatalytic reaction order. Ideas for future research on crystal facet engineering
for improving the performance of photocatalysts are also proposed.

1. Introduction This is because many applications such as heterogeneous


catalysis, gas sensing and ion detecting, molecule adsorption,
Engineering the shapes of semiconducting functional materials energy conversion and storage are very sensitive to surface
to desirable morphologies has long been actively pursued. atomic structures, which can be finely tailored by morphology
a control. From the intensive studies on morphology-controlled
Shenyang National Laboratory for Materials Science,
Institute of Metal Research, Chinese Academy of Sciences, materials in the past decades, significant advancements in this
72 Wenhua Road, Shenyang 110016, China. area have been achieved.1–29
E-mail: cheng@imr.ac.cn; Fax: +86 24 23903126; Semiconductor photocatalysts have attracted worldwide
Tel: +86 24 23971611
b attention for their potentials in environmental and energy
Department of Chemistry and Institute of Environment,
Energy and Sustainability, The Chinese University of Hong Kong, applications.30–57 The process of heterogeneous photocatalysis
Shatin, New Territories, Hong Kong, China. with semiconductor photocatalyst involves three mechanistic
E-mail: jimyu@cuhk.edu.hk; Fax: (+852) 2603-5057; steps: the excitation, bulk diffusion and surface transfer of
Tel: (+852) 2609-6268
c
ARC Centre of Excellence for Functional Nanomaterials, photoexcited electrons and holes. Apparently, the reactivity of
The University of Queensland, Qld. 4072, Australia the photocatalyst is definitely affected by surface atomic

Gang Liu received his Bachelor Jimmy C. Yu received his PhD


degree in Materials Physics in in Chemistry from University
Jilin University in 2003. He of Idaho in 1985. After teaching
obtained his PhD degree in for 10 years in the US, he
Materials Science at Institute returned to Hong Kong to
of Metal Research (IMR), establish an environmental
Chinese Academy of Sciences science programme at The
(CAS) in 2009. During his Chinese University of Hong
PhD study, he worked at Prof. Kong. He is now Professor
G. Q. Max Lu’s laboratory for and Deputy Director of the
1.5 years in Australia. Now he Institute of Environment,
is the recipient of the T. S. Keˆ Energy and Sustainability.
Research Fellowship founded With an h-index of 50, Prof.
by the Shenyang National Yu has published many highly
Gang Liu Laboratory for Materials Jimmy C. Yu cited papers on photocatalytic
Science (SYNL), IMR CAS. nanomaterials. He is also
His current research interests are to explore semiconductor active in knowledge transfer. Photocatalytic air and water treat-
photocatalysts to utilize solar energy for clean environment ment systems based on his patents are commercially available.
and renewable energy. Prof. Yu was appointed Chang Jiang Scholar Chair Professor
by the Chinese Ministry of Education in 2009.

This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 6763–6783 6763
View Article Online

structures because surface atomic arrangement and coordination The products from solar-induced photoreduction of CO2 often
intrinsically determines the adsorption of reactant molecules, consist of mixed hydrocarbons including CH4, CH3OH and
surface transfer between photoexcited electrons and reactant HCOOH.51,81–84 Such poor selectivity may be due to the fact
molecules, and desorption of product molecules. Surface atomic that most photocatalysts are terminated by too many facets
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

arrangement and coordination changes with crystal facets in resulting in excessively dispersed surface atomic and electronic
different orientations. Therefore, the reactivity of a given photo- structures. This causes different adsorption states of reactant
catalyst sensitively varies on the enclosed crystal facets. molecules, and generates photoexcited electrons of different
Although intensive experimental and theoretical studies energies on different crystal facets. Although there were
have been conducted on the reactivity of different surfaces of attempts for improving the selectivity,78,85–90 photocatalytic
metal oxides such as TiO2,44,58–63 much less attention is paid selectivity control is still in its infancy. It is expected that new
on the dependence of photocatalytic reactivity on crystal opportunities may be created by lowering the dispersivity of
facets in different orientations.64,65 Stimulated by the rapid surface atomic structure and electronic structure of photo-
development of controlling the morphologies of materials in catalysts via crystal facet engineering. A breakthrough in this
nanoscience and nanotechnology, interest has emerged in research will intensify further development in selective organic
tuning crystal facets of photocatalysts in order to optimize transformations based on semiconductor photocatalysis.77,79
the photocatalytic reactivity. For example, the morphology of Several insightful review papers have reported the experimental
anatase TiO2, one of the most important photocatalysts, has and theoretical aspects of crystal shape engineering.26,91–95
been investigated for sphere, rod, wire, tube, slightly and However, there has been no comprehensive review specially
heavily truncated octahedron, belt and sheet structures.21,27,66–70 dealing with the crystal facet engineering of semiconductor
Some of these morphologies are enclosed with recognizable photocatalysts.94,95 In this reivew, we introduce the motiva-
facets such as {001}, {101} and {010}.21,66,69 By tuning the ratio tions and basic strategies of crystal facet engineering of
photocatalysts. Most attention will focus on the advances
of different facets, the photocatalytic reactivity would be corres-
made in synthesizing faceted photocatalysts, in particular
pondingly changed. It is conventionally considered that a facet
TiO2, and the unique properties of the faceted photo-
with a high percentage of under-coordinated atoms possesses a
catalysts. This paper may shed some light on how crystal facet
superior reactivity to that with a low percentage of under-
engineering can be better applied to design more efficient
coordinated atoms.71–74 A cooperative mechanism involving
photocatalysts.
both favorable surface atomic structures and surface electronic
structures has been proposed recently to account for the
differences in reactivity.75 Unfortunately, reactive facets usually
have a relatively high surface energy. They are therefore unstable
2. Basic strategies of crystal facet engineering
during the crystal growth as a result of reducing total surface The evolution of crystal shape during growth is basically
energy of crystals. Acquiring a high percentage of reactive facets driven by continuously decreasing the total surface energy of
by crystal facet engineering is highly desirable for improving the crystal, and finally stops at the minimum surface energy point
photocatalytic reactivity. in a given growth environment. Such process is apparently
High selectivity in photocatalysis is difficult to achieve. related to the species and concentration of components in the
Many photocatalysts show high reactivity for the degrada- growth environment, which provides a flexible platform to
tion of organic molecules but usually with poor selectivity.76–80 play with the shapes of crystals. To tailor the fractions of

G. Q. Max Lu received his Hui-Ming Cheng received his


PhD in Chemical Engineering PhD degree in Materials
from University of Queensland, Science from Institute of
Australia. He worked as a Metal Research, Chinese
Lecturer in Nanyang Techno- Academy of Sciences (IMR
logical University, Singapore CAS) in 1992. He worked at
before returning to Queensland AIST (Kyushu) and Nagasaki
where he has been Senior University of Japan, and MIT,
Lecturer, Reader, Professor USA. He is Professor and
and Chair in Nanotechnology. Head of Advanced Carbon
He has received numerous Research Division in Shenyang
prestigious awards including National Laboratory for
the Federation Fellowship Materials Science, IMR CAS.
(twice). He has been Director His current research interests
Gao Qing (Max) Lu for the Australian Research Hui-Ming Cheng focus on carbon nanomaterials
Council Centre for Functional (carbon nanotubes, graphene
Nanomaterials since 2003. Professor Lu is co-author of 400 and nanoporous carbon) and new energy materials for high-
journal publications with an h-index of 48, and total citations of performance batteries, supercapacitors, and solar energy conversion.
over 9100. His research interests include carbon, silicate and
oxide nanoparticles and nanoporous materials for energy and
environmental applications.

6764 Chem. Commun., 2011, 47, 6763–6783 This journal is c The Royal Society of Chemistry 2011
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM. View Article Online

Fig. 1 Two-step nucleation of colloidal particles observed at 800 Hz


and 167 V cm 1: (a) Initial dilute liquid phase. (b) Amorphous dense
droplets are first created from the mother phase. (c) Crystalline nuclei
are created from the amorphous phase. Reprinted with permission
from ref. 99. Copyright 2007, American Chemical Society.

terminated facets, both bottom-up and top-down synthesis


Fig. 2 Schematic of the effect of solvent and additive/impurity
routes are available.
molecules or ions on the morphological control of crystal facets.
The bottom-up route for crystallization from solution involves
the nucleation and growth of crystals. According to the
classical nucleation theory, this is supposed to be a single-step mother-particles in a given solvent environment, which is also
nucleation process. However, recent studies strongly suggest a analogical to that of the Karst topography involving two
two-step nucleation, with an initially created dense liquid stages of etch and etch-stop.106 Therefore, the key in this route
nucleus and subsequent formation of a crystalline nucleus in is to choose appropriate etching and morphology-controlling
the dense liquid (see Fig. 1).96–99 The significance of the two- (capping) agents so that the purpose of growing crystals with
step nucleation is that extended space and time for affecting desired facets can be reached. The mechanism of etching agent
the morphology of crystals might provide more chances to and capping agent is that the capping agent molecules or ions
control the whole crystallization process by changing the preferentially adsorb on some surfaces to prevent them attack
reaction conditions. by etching agent while the etching agent molecules or ions will
The density, size distribution and surface chemistry of the chelate with cations on exposed surfaces and subsequently
resulting crystalline nucleus is highly dependent on the reaction cause the dissolution of the chelated surfaces. Depending on
environment of the solution. Therefore, the subsequent growth the property of the mother-crystal, the chosen appropriate
behaviours (i.e. growing rate and orientation) of surviving etching agent can vary substantially. For example, hematite
nucleus after coarsening, and thus the final shape can be quite Fe2O3 crystals with major {110} and minor {001} facets can be
different. It is well established that solvents, impurities and selectively etched by oxalic acid along the [001] direction with
additives in solution can substantially influence the ultimate phosphate ions as capping agent preferentially adsorbed on
shape of the crystals.100–104 In summary, the solvent effect on {110}, so that the resultant discs are reversely terminated with
the crystal shape is generated via the varied solvent–solute major {001} and minor {110}.105 In contrast, urea acting as
interactions along different orientations of a crystal. This is both etching and capping agent is an appropriate candidate
because surface atomic arrangement and thus surface affinity for growing WO3 octahedron crystals from irregular WO3
for the solvent to each orientation is different, which can particles.106
certainly affect the growth rates to the crystal surfaces and The above part deals with the controllable growth of single
hence the final shape of the crystal.91 Owing to the possible crystal particles with well defined facets. These faceted crystal
strong interaction between adsorbates and faces with some units can also be considered as building blocks for constructing
atomic configurations, introducing impurities or additives in a more complex structures so that the facets of crystals can be
given solution environment can further tune the growth rates selectively exposed outwards in the constructed structures. The
to the crystal faces and thus tailor the shape of the crystal by assembly mechanism of forming complex structures usually
preferentially adsorbing on some crystal faces. The mechanism involves template methods, Ostwald ripening (‘‘the growth of
of tailoring the crystal shape by impurity/additive effect is to larger crystals from those of smaller size which have a higher
change surface relative stability of facets in terms of surface solubility than the larger ones’’—IUPAC Compendium of
energy. The effectiveness of this effect depends closely on the Chemical Terminology, 2nd edition, 1977), Kirkendall type
match of impurity and additive with faces in a given solution diffusion (normally referring to comparative diffusive migrations
environment. The role of solvent and impurity/additive effect among different atomic species in metals and alloys under
in controlling the shape of the crystal is schematically illuminated heating conditions),28 oriented attachment (a combination
in Fig. 2. of crystallites through their suitable surface planes),28 space-
Besides the widely employed bottom-up route, the top-down predefined growth, and probably their combination, which have
route can also be used to control the facets of single crystal been fully discussed by Zeng.28 In most cases, the constructed
particles. So far, this approach has been adopted in several structures are polycrystalline. However, single crystalline nano-
systems such as Fe2O3,105 WO3106 and TiO2.107 The nature of wires have been realized in the case of the assembly of CdTe
this route is the partial dissolution of mother-crystal particles nanoparticles driven by strong dipole–dipole interaction.108
and/or subsequent recrystallization on the surfaces of residual Recently, Schliehe et al. reported single crystalline ultrathin

This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 6763–6783 6765
View Article Online

PbS sheets by two-dimensional oriented attachment of PbS predicted to be as high as 94%. Although the surface energy
nanocrystals.109 of {010} (0.53 J m 2) was calculated to be between {001}
(0.90 J m 2) and {101} (0.44 J m 2),111 it is surprising that
no {010} will appear in the equilibrium shape of anatase.
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

3. State of the art of crystal facet engineering Regarding rutile, the predicted equilibrium shape of a macro-
In this section, we first briefly present the definition of surface scopic crystal was constructed with (110), (100), (001) and
energy and Wulff construction, which are the most basic (011) faces (see Fig. 3).112 It is found that the most stable (110)
parameters and method for predicting the equilibrium morpho- face with the lowest surface energy of 15.6 meV au 2 dominates
logy of a crystal. Detailed description can be found in the the shape while (001) with the highest surface energy of
review paper by Lovette et al.91 Crystalline TiO2, the most 28.9 meV au 2 does not exist in the equilibrium shape at all.
intensively investigated semiconductor photocatalyst, is chosen Gong et al. presented the comprehensive results of the struc-
as the most representative example to indicate the develop- tures and energetics of ten stoichiometric 1  1 low-index
ment and resultant unique properties of photocatalysts with surfaces with different possible terminations of brookite.113
well defined facets. Meanwhile, other photocatalysts are The relative stabilities of different faces determined are found
briefly introduced. to be negatively related to the concentration of exposed
coordinatively unsaturated Ti atoms. Fig. 3 shows the equilibrium
3.1 The surface energy and Wulff construction shape of brookite crystal, most of which is composed of (111),
(210), (010) and reconstructed (001) facets. It is interesting to
The energy of the terminated surface of a solid material is
note that brookite (210) as one of the most stable facets has
always higher than that of the bulk and this energy difference
very similar atomic structure to the most stable facet (101) of
is defined as the surface energy. Based on surface energies of
anatase. However, they show differences in electronic states,
all facets, one can use the Wulff construction to determine the
which may result in different chemical reactivities.113
equilibrium morphology of a material. In the Wulff construc-
All the above equilibrium shapes predicted of TiO2 poly-
tion, surface energy minimization is the central standard to
morph crystals are conducted in vacuum at absolute zero
optimize the composition of the crystal surface.
temperature, which is clearly different from the actual condi-
3.2 Predicted shapes of TiO2 polymorph crystals tions. Barnard et al. found that the presence of water has
influence on the shapes by changing the size of anatase {001}
Among various photocatalysts, it is little doubt that crystalline and aspect ratio of rutile.114 Therefore, these predicted shapes
TiO2 as the most intensive studied photocatalyst plays an are definitely derived from the observed morphologies of
indispensible role in promoting the development of photocatalysis natural or synthetic crystals, to some extent, influenced by
since the discovery of TiO2 photoanode for photocatalytic the environmental growth conditions. Despite some differences
hydrogen evolution in 1972.30 TiO2 polymorphs comprise between the vacuum conditions used in the calculations and
anatase, rutile and brookite. It is generally considered that the real conditions for crystal growth, the most stable facets
anatase is the most active, rutile is less active, and brookite predicted usually show the largest fraction in crystal surfaces.
is not active at all for photocatalysis applications, though all In the practical synthesis procedures, TiO2 crystals are
three phases are constructed with TiO6 octahedra. Anatase always subjected to either acidic or alkaline conditions, and
and rutile are tetragonal and brookite is rhombohedral. Due therefore their morphologies varies with the surface chemistry
to different arrangements of these TiO6 octahedra in the involved. Based on surface free energies and surface tensions
polymorphs, the surface terminations in various orientations from first-principles calculations, Barnard et al.115 systemati-
and thus the equilibrium morphologies of the crystals are cally studied the effects of surface chemistry in terms of acidic
completely different. and alkaline conditions on the morphologies of anatase and
According to the Wulff construction and calculated surface rutile nanoparticles. As shown in Fig. 4a–e, the surface
energy, the shape of anatase under equilibrium conditions is a termination by hydrogen (acidic conditions) results in little
slightly truncated tetragonal bipyramid enclosed with the eight change in the shapes of both polymorphs relative to vacuum.
isosceles trapezoidal surfaces of {101} and two top squares of However, in water terminated surfaces and hydrogen-poor
{001}, as shown in Fig. 3.110 The percentage of {101} is surfaces, in particular oxygenated surfaces, both polymorphs
are apparently elongated. As a result, it is noteworthy that new
facets {010} as the ‘‘belt’’ appear at the center of anatase
particles. These predictions are very important to experi-
mentally realize the fine tuning of morphology by controlling
surface chemistry.
Retaining a large percentage of facets with high surface
energies (i.e. anatase {001}) in crystals usually result from their
rapid growth rates during crystal growth. Appropriate surface
adsorbates may substantially lower the surface energy and
Fig. 3 The equilibrium shape of a TiO2 crystal in the anatase, rutile thus probably obtain a high percentage of high-energy surfaces.
and brookite, according to the Wulff construction and the calculated Yang et al.21 started the theoretical investigation the influence
surface energies. Reprinted with permission from refs. 110, 112 and of surface terminated non-mental atoms on the surface energies
113. Copyright 2001, 1994, 2007, American Physical Society. of anatase (101) and (001), and found that surface terminated

6766 Chem. Commun., 2011, 47, 6763–6783 This journal is c The Royal Society of Chemistry 2011
View Article Online

3.3 Anatase TiO2 crystals with well-defined facets


3.3.1 Anatase with a large percentage of {101} facets. As
indicated above, {101} facets will dominate the surface of most
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

anatase particles prepared conventionally. Well faceted {101}


facets can be achieved by coarsening, in terms of the growth of
large particles at the cost of small particles (namely, Ostwald
ripening). As early as 1999, slightly truncated octahedron
crystals with well defined {101} facets were obtained by
coarsening pristine titania particles from sol–gel route under
various hydrothermal conditions.66 Acidic conditions were
found to accelerate the coarsening relative to deionized water.
The percentage of {101} varies as a result of changed growth
Fig. 4 Morphology predicted for anatase (top) with (a) hydrogenated rates along o0014 and o1014 directions during different
surfaces, (b) with hydrogen-rich surface adsorbates, (c) hydrated growth stages. The driving force for simple particle growth is
surfaces, (d) hydrogen-poor adsorbates and (e) oxygenated surfaces, the reduction in surface energy. Minimizing the area of
and rutile (bottom) with (f) hydrogenated surfaces, (g) with hydrogen- high surface energy faces to further reduce surface energy is
rich surface adsorbates, (h) hydrated surfaces, (i) hydrogen-poor responsible for morphology evolution.
adsorbates and (j) oxygenated surfaces. Reprinted with permission
Recently, impressive approaches have been reported to
from ref. 115. Copyright 2005, American Chemical Society.
obtain anatase with a high percentage of {101} with the rapid
development of morphological control. Amano et al.116 reported
fluorine could effectively reverse the stabilities of (101) and
a hydrothermal transformation directly from K-titanate nano-
(001). The maximum percentage of {001} was predicted to be
wires (5–15 nm in diameter and several hundreds of nanometres
more than 90% (See Fig. 5). Their subsequent experimental
in length) to octahedral anatase single crystals (particle size is
results solidly confirmed the key role of surface fluorine in
less than 100 nm) with predominant {101} as shown in Fig. 6.
stabilizing {001}. This work has triggered the subsequent
Dai et al.117 employed very small particles of amorphous TiO2,
intensive research interests in preparing faceted anatase and
prepared by electrospinning, as precursor to prepare truncated
also other photocatalysts.
tetragonal bipyramids of anatase in size of less than 50 nm via a
hydrothermal route. The percentage of {101} can be as high
as 90%.
All the above synthesis procedures have a common feature
of using pre-prepared less crystalline TiO2 particles or nano-
structured titanates as precursor. Different from these routes,
Liu et al.118 developed a novel route to prepare TiO2 with
crystalline bulk TiB2 as precursor via an acidic hydrothermal
route. Due to its specific bonds and intrinsic chemical stability
together with its ability for controlled release of titanium
species, TiB2 shows very powerful versatility in tuning both
crystal phases and morphologies of TiO2 by simply controlling

Fig. 5 Slab models and calculated surface energies of anatase TiO2


(001) and (101) surfaces. (a, b) Unrelaxed, clean (001) and (101)
surfaces. Ti and O atoms are represented by grey and red spheres,
with six-fold Ti, five-fold Ti, three-fold O and two-fold O labelled as
6c-Ti, 5c-Ti, 3c-O and 2c-O, respectively. (c, d) Unrelaxed (001) and
(101) surfaces surrounded by adsorbate X atoms. (e) Calculated
energies of the (001) and (101) surfaces surrounded by X atoms. Fig. 6 SEM image of (a) titanate nanowires and (b, c) particles after
(f) Plots of the optimized value of B/A and percentage of {001} facets hydrothermal reaction of titanate nanowires. (d) TEM image and
for anatase single crystals with various adsorbate atoms X.21 electron diffraction pattern of an octahedral bipyramid.116

This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 6763–6783 6767
View Article Online

the species and amount of mineralizers (Cl , SO42 and since Yang et al.21 realized the first case of 47% {001} in
NO3 ). Specifically, anatase bipyramids with a high percen- uniform crystals. To this end, improved synthesis procedures
tage of {101} can be obtained by tuning the ratio of SO42 to have been developed and exciting advances have been
Cl in solution. This new strategy could also be extendable to obtained as summarized in Table 1.117,120–139
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

other metal borides such as ZrB2, TaB, Nb3B4 to synthesize The percentage of {001} facets can be well tuned by
their corresponding metal oxides. controlling the concentrations of precursors and fluorine
Although the percentage of {101} in the synthesized species, also the synergistic effects of other capping agents.
anatase bipyramids can be nearly 100%, it is impossible to For instance, the percentage can increase from 18 to 72% by
completely eliminate high-energy {001}. This is balanced by simply decreasing the concentration of Ti(SO4)2 from 100 to
the minimization required for the total surface energy of whole 10 mM.129 Zhang et al.140 showed that the percentage gradually
particle. If {001} grew out of existence, the intersection of improved with increasing amount of 1-butyl-3-methylimidazolium
{101} planes would result in an acute angle, which would tetrafluoroborate. The reason for the improved percentage is
presumably show higher energy than the high energy (001) the increased surface coverage fractions of fluorine. It is
surface.66 theoretically predicted that the different surface coverage
To this end, a large percentage of {101} facets is usually fractions can greatly affect the surface energy and thus the
achieved in bipyramids, which is the equilibrium shape of percentage of {001} during the growth of crystals.129 The
anatase. With respect to the bipyramids, anatase nanobelts calculated surface energies decrease monotonously from
with two large (101) faces have been recently realized by Wu 0.96 to 0.64 J m 2 with increasing coverage. On the other
et al.119 The synthesis strategy is to hydrothermally treat hand, by employing alcohol as synergistic capping agent, the
titanium dioxide powder in NaOH solution and the resultant percentage can be further improved: 2-propanol, ethanol,
products after washing with hydrochloric acid were thermally tert-butanol and benzyl alcohol are experimentally proved
treated to obtain the final belts (Fig. 7). Experimental to be effective in obtaining a higher percentage of {001}
and theoretical studies have revealed that the belt structure facets.121,141 Yang et al.121 theoretically revealed that the
has obvious merits in lowering the recombination of adsorption energy of fluorine on (001) surface increased by
photoexcited electrons and holes, in contrast to nanospheres, 0.8 eV in contrast to the slightly weakened adsorption energy
as a result of greater charge mobility, fewer localized states by 0.1 eV on (101) when the derived (CH3)2CHO from
near the band edges, and effective trapping of oxygen on the 2-propanol was bound on surface unsaturated Ti atoms. The
(101) face. maximum percentage of {001} could reach 89% with the
involvement of only fluorine.120
3.3.2 Anatase TiO2 with a large percentage of {001} facets. It is apparent that the substantial role of fluorine in stabilizing
It is no doubt that a large percentage of {001} facets has become {001} facets has been extensively demonstrated in various
a very popular target in the synthesis of anatase TiO2 crystals aqueous synthesis routes, though the fluorine may be released
from different sources. So far, the sources of fluorine includes
the most frequently used hydrofluoric acid (HF),21 1-butyl-3-
methylimidazolium tetrafluoroborate,123 ammonium bifluoride
(NH4HF2),142 ammonium fluoride (NH4F),125 and titanium
tetrafluoride (TiF4) precursor itself.134 No evidence suggests
the dependence of the percentage of {001} on the source of
fluorine.
Regarding the precursors for TiO2, the dissolvable TiF4,
Ti(SO4)2 and Ti(OBu)4 have been commonly used to prepare
anatase with {001} facets. It seems that all precursors can be
used for tailoring both the percentage of {001} facets and
particle size. Therefore, the type of dissolvable precursors is
not a decisive factor in growing the desired pure anatase
crystals. However, some undissolvable precursors such as
Ti,130,139 TiN,126 TiS2133 and TiB2128 crystalline powder can
play an important role in securing some of the unique properties
including visible light absorption and oxygen deficiency. In
addition, TiO2 films with oriented anatase {001} facets can be
Fig. 7 (a) SEM images of anatase TiO2 nanobelts; (b) bright-field directly grown on Ti foil substrates by a simple hydrothermal
TEM image of a TiO2 nanobelt; the inset is a selected area electron treatment of Ti foils in HF solution.143
diffraction (SAED) pattern taken along the [100] direction of the
Besides the commonly-employed wet-chemistry routes, two
nanobelt; (c) HRTEM image of a nanobelt; (d) bright-field TEM
other methods have also been developed for preparing anatase
images and corresponding SAED patterns of a TiO2 nanobelt recorded
in different crystallographic orientations while keeping the [010] axis crystals with a high percentage of {001} facets of good quality.
horizontal, as well as schematic illustrations of the nanobelt and the One is to directly convert amorphous TiO2 nanotube arrays on
relationship between the incident beam and the nanobelt while taking Ti foil to anatase crystals enclosed with well defined {101} and
images and diffraction patterns. Reprinted with permission from {001} by calcining at a temperature above 400 1C in air for
ref. 119. Copyright 2010, American Chemical Society. 2 h.125 Furthermore, both the particle size and the facet

6768 Chem. Commun., 2011, 47, 6763–6783 This journal is c The Royal Society of Chemistry 2011
View Article Online

Table 1 Summary of typical SEM/TEM images, particle size, percentage of 001 facets, photocatalytic activity of anatase TiO2 crystals with
preferential {001} facets from different synthesis procedures. Reprinted with permission from ref. 21, Copyright 2008, Nature Publishing Group;
Reprinted with permission from ref. 120, 121, 124, 125, 126, 128, 131, 135, 140, Copyrights 2009, 2009, 2009, 2009, 2009, 2009, 2010, 2010, 2010,
American Chemical Society. Reprinted with permission from ref. 127, Copyright 2009, Wiley-VCH; Reprinted with permission from ref. 133.
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

Copyright 2010, Elsevier

Particle size and


SEM/TEM Image percentage of {001} Photocatalytic activity Typical synthesis parameters Ref.
Size: ca. 1.66 mm No data Precursor: 30 mL 5.33 mM TiF4 21
Percentage: 47% Morphology controlling agent: 0.4 mL
10 wt% HF
Procedure: hydrothermal process, at 180 1C,
414 h

Size: ca. 1.09 mm 5 times higher photoreactivity normalized by Precursor: 14.5 mL 2.76 mM TiF4 121
Percentage: 64% surface area in generating  OH radicals than Morphology controlling agent: 0.5 mL
P25 TiO2. 10 wt% HF, 13.38 mL 2-propanol
Procedure: hydrothermal process,
at 180 1C, 411 h

Size: ca. 2 mm Exhibiting higher reactivity in photodegrada- Precursor: 30 mL 20 mM TiF4 123


Percentage: maximum tion of 4-chlorophenol than selected anatase Morphology controlling agent: 1-butyl-3-
80% TiO2 without dominant {001} facets. methylimidazolium tetrafluoroborate (1 mL)
Procedure: microwave-assisted hydrothermal
process, at 210 1C, 90 min

Size: ca. 2.2 mm Both the NO removal rate and the Precursor: 30 mL 40 mM TiF4 140
Percentage: 27–50% photoactivity in decomposing aqueous Morphology controlling agent: ionic liquid
4-chlorophenol increase with increasing 1-methylimidazolium tetrafluoroborate
percentage of {001} facets from 0% to 50%. (0.5, 1 or 2 mL)
Procedure: microwave-assisted hydrothermal
process, at 210 1C, 90 min

Size: ca. 2 mm Improved photoreactivity in decomposing Precursor: 10 mg Ti powder/30 mL aqueous 130


Percentage: no data methylene blue from 48.1 to 82.5% relative solution containing 3 mL H2O2
to irregular TiO2. Morphology controlling agent: 0.1 mL HF
(40 wt%)
Procedure: hydrothermal process, at 180 1C,
10 h

Size: ca. 50–250 nm H2 evolution: showing a higher rate than P25 Precursor: TiCl4 vapor generated by bubbling
Percentage: TiO2, in particular in the presence of argon (200 mL min 1) into TiCl4 solution at 124
B/A = 0.65–0.75 [Pt(NH3)4]Cl2 under 400 W high-pressure 358 K mixing O2 (1200 mL min 1)
mercury lamp; Procedure: CVD, 1573 K.
Photodecomposition capability: as high as
that of P25 for decomposing acetic acid, much
higher than P25 for decomposing methanol.

Precursor: amorphous TiO2 nanotube arrays


Size: decreasing from 220 No data
with a length 430 mm from the anodization
to 20 nm with the increase 125
process of Ti foil in electrolyte containing
of NH4F from 0.1 to 2%
NH4F
Percentage: Procedure: heat treatment at 4 400 1C for 2 h
B/A = 0.55–0.82 with a ramping rate 416 1C min 1 in air, two
sides of TiO2 tube arrays are covered by two
glass plates.

Size: ca. 30–85 nm Exhibiting photoreactivity both for generating Precursor: 100 mM Ti(SO4)2 129

Percentage: 18% OH radicals and for hydrogen evolution that Morphology controlling agent: 0.2 M 40 mL
is markedly superior—by factors of 5.6 and HF
8.2, respectively—to ca. 3 mm anatase with Procedure: hydrothermal process, at 180 1C,
72% {001} facets. 22 h

Size: ca. 30–130 nm In contrast to the substantially prolonged Precursor: tetrabutyl titanate, 5 mL Ti(OBu)4 120
Percentage: 68–89% degradation half-rate of methylene orange Morphology controlling agent: 0.4–0.8 mL
with P25 TiO2, the samples obtained exhibit a 47 wt% HF
gradually shortened half-rate in 7 cycles. Procedure: hydrothermal process,
at 180/200 1C, 24 h

This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 6763–6783 6769
View Article Online

Table 1 (continued )

Particle size and


SEM/TEM Image percentage of {001} Photocatalytic activity Typical synthesis parameters Ref.
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

Size: ca. 23–94 nm Conversion of toluene: highest value of 28% Precursor: 0.16 g TiF4 141
Percentage: 50–65% compared to 10% for P25. Morphology controlling agent: 40 mL
Selectivity to benzaldehyde: nearly the same tert-butanol/ benzyl alcohol and tert-butanol
as that of P25 at around 90%. Procedure: hydrothermal process,
Yield of benzaldehyde: highest value of 26% at 160 1C, 72 h
compared to 9% for P25.

Size: ca. 1.09 mm Nitrogen heteroatom doping leading to a Precursor: 150 mM TiN 126
Percentage: 64% capability of generating  OH radicals and Morphology controlling agent: 30 mL 1M HF
photocatalytic hydrogen evolution under Procedure: hydrothermal process, at 180 1C,
visible light. 30 h

Size: ca. 1.2 mm Sulfur doping causing the capability of Precursor: 9 mM TiS2 133
Percentage: 41% generating  OH radicals and decomposing Morphology controlling agent: 20 mL 50 mM
Rhodamine B under visible light. HF
Procedure: hydrothermal process, at 180 1C,
12 h

Size: ca. 4 mm Oxygen deficiency greatly improving Precursor: 100 mM TiB2 128
Percentage: 56% photocatalytic hydrogen evolution. Morphology controlling agent: 20 mL 1 M
HF
Procedure: hydrothermal process at 180 1C,
14 h

Size: ca. 1 mm No data Precursor: 6 mL 0.04 M TiF4 134


Percentage: 22.8% Morphology controlling agent: disodium
ethylenediaminetetraacetate (EDTA, 6 mM)
Procedure: hydrothermal process at 200 1C,
4h

Size: ca. 130 nm Photoelectrochemical water splitting Precursor: Ti foil 143


Percentage: patterned Morphology controlling agent: HF (10 mM)
{001} outwards Procedure: hydrothermal process at 140 1C,
10 h

Size: 3–6 mm spheres Methylene orange (MO) in solution Precursor: 10 mL Ti(OBu)4 in 40 mL absolute
consisting of sheets with (20 mg L 1) can be fully decomposed by the ethanol
127
a side length and thickness sample within 10 min, while only 29% MO is Morphology controlling agent: HF
of 40 nm and 4 nm removed with P25 Procedure: hydrothermal process, at 180 1C,
Percentage: 83% 5.5 h

Size: ca. 1–2 mm spheres Depending on the degree of surface fluorine Precursor: 15 mmol Ti(SO4)2
consisting of anatase termination, the preferential decomposition of Morphology controlling agent: 15 mmol 135
polyhedra methylene blue or orange can be tuned. NH4F and ethanol
Percentage: 20% Procedure: hydrothermal process, at 180 1C,
12-24 h

Size: ca. 1.2 mm spheres No data Precursor: titanium(IV) isopropoxide


consisting of sheets with a Morphology controlling agent: 0.03 mL
131
thickness of 3 nm and size diethylenetriamine (DETA) in 42 mL
of 100–300 nm of isopropyl alcohol (IPA)
Percentage: 95–99% Procedure: hydrothermal process at 200 1C,
24 h, the subsequent calculation at 400 1C
for 2 h in air

percentage can be tuned by controlling the concentration of a gas-phase reaction of TiCl4 with O2 at a temperature of
NH4F solution for pre-soaking of TiO2 before calcination. 1300 1C.124 The most important feature is that small TiO2
The critical parameters include the temperature and its particles of 50–250 nm could be obtained at such high
ramping rate, and the presence of NH4F. The other is to use temperature without using fluorine containing reactants. It is

6770 Chem. Commun., 2011, 47, 6763–6783 This journal is c The Royal Society of Chemistry 2011
View Article Online

claimed that uniform and rapid heating at a high temperature shapes.69 The key feature of their approach is the use of water
would enable homogeneous nucleation and subsequent vapor as hydrolysis agent to accelerate the reaction and the use
growth to well faceted crystals with few defects. The low of both oleic acid and oleylamine as two distinct capping
concentration of TiCl4 and the narrow heating zone would surfactants which have different binding strengths to control
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

prevent formation of large particles and polycrystalline aggre- the growth of the TiO2 nanoparticles.69 As a result, different
gates with grain boundaries.124 shapes such as rhombic, truncated rhombic, spherical,
Can anatase crystals with a high percentage of {001} grow in dog-bone, truncated and elongated rhombic and bars could
a fluorine-free wet-chemistry route? The answer is yes. be formed, as shown in Fig. 8. Although the authors
Chen et al. proved that diethylenetriamine (DETA) in did not identify crystallographic facets of the exposed
isopropyl alcohol could act as an alternative morphology surface, it is easy to conclude that compared to untruncated/
controlling agent to fluorine species for stabilizing high energy truncated rhombic crystals, the elongated crystals are
(001) faces in the solvothermal system.131 As-prepared TiO2 enclosed with an additional four {010} facets based on the
microspheres consisting of {001} dominant sheets are poor in symmetries and predicted morphologies of anatase. However,
crystallinity. Post-calcination at 400 1C in air leads to heavy coating of surface by surfactants makes it impossible
increased crystallinity without morphology change. The to evaluate the photoreactivity of the crystals with different
percentage of {001} is claimed as nearly 100%. Then, a further shapes.
challenge is whether or not anatase crystals with a high Very recently, Li et al. used Na-titanate nanotubes as
percentage of {001} can be grown without both fluorine and precursor to prepare uniform anatase tetragonal faceted-
other capping agents. Self-assembly of nanosized building nanorods with a large percentage of {010} facets by a hydro-
blocks with {001} into micron-sized hierarchical structures is thermal process in basic solution (see Fig. 9).146 The proposed
also studied.127 The obvious advantage of this method is the growth mechanism involves the formation of Ti(OH)4 frag-
stabilized ultrathin TiO2 structures and facile separation from ments from titanate nanotubes and subsequent anatase crystal
solution. nuclei by dehydration reaction between Ti–OH and HO–Ti. It
Zhou et al.144 proposed the interesting oriented transforma- is considered that the continuous release of OH during the
tion from micron-sized NH4TiOF3 square plates consisting of hydrothermal process plays a key role in generating {010}
mesocrystals to anatase TiO2 without changing the original facets. Barnard et al.115 revealed that O-terminated (010) had a
architecture due to the well matching lattices of {001} and lower surface energy than O-terminated (101) and (001). It was
{010}/{100} facets in these two materials. Detailed structure
analysis shows that the obtained anatase crystals are enclosed
with {001} and {010}/{100}.
From the energy saving view point, it is very desirable to
shorten the duration of crystal growth as much as possible
before a good percentage of {001} facets can be reached.
Zhang et al. has achieved a good step towards this goal.123
With the assistance of microwave radiation, the period of
hydrothermal process has been substantially reduced from the
usual 410 h to 90 min, while the percentage of {001} can be
kept as high as ca. 80% in the well defined micron-sized
crystals. Further efforts in this direction may lead to new
developments of nanosized anatase with a higher percentage
of {001} facets.

3.3.3 Anatase with a large percentage of {010} facets.


Compared to the widely investigated {101} and {001}, low-
index {010} facets are experimentally less studied until
recently. Actually, as far as the density of surface unsaturated
Ti atoms is concerned, {010} has the same 100% Ti5c as {001}.
According to the commonly-used criteria to predict the
reactivity of surface, {010} is also favourable for hetero-
geneous reactions. Furthermore, theoretical results have
predicted that, contrary to water being adsorbed molecularly
on (101) face and similar to that water partially dissociates on
Fig. 8 Schematic illustration of the overall formation and shape
the (001) face, water molecules can be adsorbed dissociatively
evolution of TiO2 NCs. TEM images of (A) dog-bone-shaped TiO2
on (010) face. Acquiring a high percentage of {010} facets is
obtained at TB/OA/OM (TB: titanium n-butoxide; OA: oleic acid;
therefore of great significance in optimizing the photocatalytic OM: oleylamine.) = 2 : 6 : 4; (B) truncated and elongated rhombic
activity of anatase. TiO2 obtained at TB/OA/OM = 2 : 5 : 5. Insets are high-magnification
Fabrication of anatase crystals with 43.4% of {010} facets images of the corresponding shapes (left, longitudinal view; right,
was reported by Wu et al. in 2008.145 In 2009, Dinh et al. cross view on panel B). Reprinted with permission from ref. 69.
reported powerful routes to prepare anatase crystals in various Copyright 2009, American Chemical Society.

This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 6763–6783 6771
View Article Online

3.4 Rutile TiO2 with defined facets


Rutile is the thermodynamically stable phase of TiO2 poly-
morphs. It can be obtained by either the hydrolysis of Ti
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

precursor and subsequent crystallization or the post-


transformation from anatase/brookite phase via thermal treat-
ment (phase transformation temperature required depends on
the particle size of TiO2)148 and mechanical processing.149
Although rutile is considered to be less active in photocatalytic
reactions with respect to anatase, nanostructured rutile has
also been used in photocatalysis applications and some cases
Fig. 9 (a) Low- and (b) high-magnification SEM images of anatase
show higher activity than anatase. One obvious merit is its
TiO2 tetragonal faceted-nanorods (TFNR). The inset of (b) is the top smaller bandgap than anatase by 0.2 eV and thus a wider
view of a single TFNR.146 absorption range.
Various morphologies of rutile have been obtained.27,150–156
The frequently observed morphology of nanostructured rutile
believed that the preferential adsorption of hydroxyl ion on
appears as bundled or self-assembled nanowire/nanorods, as a
(010) face contributed to stabilize the (010) face and hence the
result of the preferential growth along [001] direction, as
high percentage. The above route is similar to that for anatase
shown in Fig. 11.152 Based on the crystallographic symmetries
octahedra with {001}, where nanowires instead of nanotubes
and predicted equilibrium shape of rutile, the exposed lateral
are used as precursor.
surface of these nanowires/nanorods which are parallel to the
By controlling carefully the preparation parameters
[001] direction, must be {110} facets. The ends usually have
(concentrations of precursor and morphological controlling
two types of facets: a square surface perpendicular to [001]
agent and reaction time) to satisfy the surface energy require-
direction and four triangle surfaces at an angle of around 1301
ment for each facet, Pan et al. prepared anatase single crystals
relative to {110}. The square face is {001} and the triangle face
with a respective predominance of {001}, {101} and {010} as
is {111}.
shown in Fig. 10. These products made it possible to compare
The synthesis routes of such rutile nanorods with a
the relative photocatalytic activity of each facet and then
high aspect ratio have been well documented in the
determine the photoreactivity order of facets.
literature.151,155,157–166 Generally speaking, the presence of
An alternative route via topotactic transformation reaction
Cl ions as mineralizer in the synthesis system is favourable
from layered titanate nanosheets has also been developed to
for rutile TiO2, regardless of the source of Cl . This is because
obtain anatase with preferential exposure of {010}.147
that due to different coordination ability and spatial steric
Different from the conventional hydrothermal process
effects of the mineralizers, the crystal phase of TiO2 can be
involving the nucleation and growth of anatase crystals, the
tailored through changing the mineralizers, which affect the
TiO6 octahedra in titanate nanosheets slightly and equably
linkage of six-fold coordinated monomers in different bonding
shift from the positions of the layered lattice to the positions of
modes. According to ligand field theory,156 the nucleation and
the anatase lattice in the topotactic transformation reaction.
crystallization of TiO2 in either anatase, rutile or brookite
This method gives important insight for preparing other metal
phase are strongly affected by the ligands and dehydration
oxides with preferential facets.

Fig. 10 (a) Schematic of anatase TiO2 with different percentages of


{101}, {001} and {010} facets; (b)–(d): scanning electron microscopy
(SEM) images of anatase crystals synthesized with different hydro-
fluoric acid (HF) aqueous solutions (120, 80 and 40 mM) containing
different amounts of TiOSO4 precursor (64, 32 and 32 mg) at 180 1C Fig. 11 XRD pattern (a), SEM (b), TEM (c) and HRTEM (d) image
for different times (12, 12 and 2 h). Reprinted with permission from of self-assembled rutile TiO2 by microwave heating TiCl3 at 200 1C for
ref. 75. Copyright 2011, Wiley-VCH. 1 min.152

6772 Chem. Commun., 2011, 47, 6763–6783 This journal is c The Royal Society of Chemistry 2011
View Article Online

reactions between possible partially hydrolysed Ti(IV) (3d0)


complexes existing in the solution such as (Ti(OH)n(A1)m(A2)o2
(A: anion, n + m + o = 6). An example demonstrating the
effectiveness of mineralizers (Cl , SO42 , NO3 ) in tailoring
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

phases of TiO2 has been given by Liu et al.118


Regarding the specific synthesis routes of controlling
morphology of rutile, two representative examples are given
to illuminate the formation of faceted rutile crystals. One is the
rapid formation of self-assembled microspheres with rutile
nanorods by microwave heating TiCl3 at 200 1C for only
1 min (see Fig. 11).152 The nanorods are exposed with {110}
and {111} facets, though the surface is not smooth probably
due to the extremely rapid growth rate. Interestingly, the
synthetic rutile nanorods have a smaller bandgap of 2.8 eV
than the conventional 3.0 eV, which allows it to work reason-
ably well under visible-light irradiation. The other is the Fig. 13 (a1–d1) Top- and (a2–d2) side-view of samples a–d obtained
reported dependence of degree of perfection of facets on by growing TiO2 on Ta foil at 180 1C at (a1, a2) 3.0, (b1, b2) 10,
(c1, c2) 15, and (d1, d2) 20 h, respectively. (b3) HAADF-STEM and
hydrothermal temperature by Kakiuchi et al.,153 where TiCl3
(b4–b6) elemental mapping images of an individual sample b
was also used as precursor together with NaCl additive. At
(scale bar = 500 nm). Reprinted with permission from ref. 154.
80 1C, only needle-like nanorods without well-recognized Copyright 2009, American Chemical Society.
facets were formed. By elevating the temperature to 200 1C,
well-developed lateral {110} and top {111} facets can be non-recognizable. In contrast, Buonsanti et al.172 developed
observed, as shown in Fig. 12. Apparently, a higher temperature a nonhydrolytic synthesis route to prepare high-quality
is favorable for growing crystals with well-developed facets. anisotropically shaped brookite nanorods with length of
Besides the widely used strategy of employing organic 30–200 nm. These rods are determined to be dominantly
surfactants or inorganic ions as capping agents on crystallo- enclosed with the longitudinal {210}/{100}, and basal {001},
graphic faces, doping into crystal lattices is also an effective which is consistent with the equilibrium shape of brookite
approach to manipulating crystal growth behavior and thus crystals predicted from the Wulff construction.
the final morphology of crystals. The basis of this approach is
both the changed surface energies and resultant crystalline 3.6 Other semiconductor photocatalysts with defined facets
lattice microstrains from the cell distortion.167 Very recently,
As early as 2006, Jang et al.65 have investigated the close
Yang et al. reported that Ta doping can cause the anisotropic
dependence of photocatalytic activity on ZnO facets. Largely
growth along [110] direction and cross-medal shape of rutile
TiO2 with dominant {001} facets (Fig. 13).154 inspired by the rapid development of faceted TiO2, engineering
crystal facets has rapidly moved to other photo-
3.5 Brookite TiO2 with defined facets catalysts including both binary and ternary systems. So far,
ZnO,174,175 Cu2O,176,177 SnO2,25 WO3,106 BiVO4173,178 and
Compared to anatase and rutile, brookite has attracted much
Zn2GeO4179 photocatalysts with well defined facets have been
less interest due to the generally considered lack of photo-
reported. For instance, ZnO nanodisks (see Fig. 14a) with a
catalytic activity. However, increasing examples have shown
large percentage of polar (0001) facets and ZnO nanowires
that brookite is also photocatalytically active and even
has unique photocatalytic properties in some cases.168–171
Among the synthetic brookites, crystal facets are usually

Fig. 12 SEM image of the rutile TiO2 deposits on silica glass Fig. 14 SEM images of ZnO nanodisks (a) and nanowires (b)
substrate. Reprinted with permission from ref. 153. Copyright 2006, (reprinted with permission from ref. 174. Copyright 2009, Elsevier),
Elsevier. WO3 octahedra crystals106 (c) and BiVO4 nanoplates173 (d).

This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 6763–6783 6773
View Article Online

with dominant {1010} facets (Fig. 14b) can be readily prepared long-term stability of the photocatalyst. The potential merits
by simply tuning the synthetic parameters.174 In contrast to of employing units with well developed facets for hetero-
most bottom-up routes, Zhao et al. developed an interesting structures are not only a more intimate contact facilitated by
top-down strategy to prepare WO3 octahedra enclosed with the flat facets but also probably generate some new
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

{111} facets from irregular WO3 particles by employing urea photocatalytic properties by the unique interface atomic
as both etching and capping agent in a simple solvothermal structures. For instance, Wang et al. reported that ZnO–CdS
route.106 The top-down approach was also used to prepare nanoparticle heterostructures could work well in a direct
Fe2O3 nanodisks with exposed {110} and {001} facets from Z-scheme (a type of charge transfer process by recombining
pristine Fe2O3 particles with oxalic acid as etching agent and low-energy photoexcited electrons and holes in the inter-
phosphate ions as capping agent.105 Monoclinic BiVO4 nano- face37), but suffered from a serious decay in photocatalytic
plates with exposed {001} facets (see Fig. 14d) have been hydrogen evolution rate by more than 50% from Na2S/Na2SO3
synthesized via a hydrothermal route without the use of any solution for 30 h reaction.185 However, by employing ZnO rods
template or organic surfactant.173 New substantial develop- enclosed with well-developed {1100} and {0001} facets to
ment of various photocatalysts with well defined facets is construct ZnO–CdS heterostructures (see Fig. 15), the hydrogen
greatly anticipated. evolution rate can retain ca. 96% of the initial rate after 30 h
reaction.186 More interestingly, Liu and co-workers discovered
3.7 Heterostructures with defined facets the first case of quite stable overall water-splitting with CdTe
Heterostructuring with different phases or materials is one of quantum dots (QDs) sensitized well-faceted ZnO nanowire
the most important strategies to develop highly efficient array photoanodes in Na2SO4 solution.187
powder photocatalysts and photoanodes. The advantages of In the above cases, CdS nanoparticles or CdTe QDs were
heterostructuring include extended light responsive range deposited on ZnO surfaces by post-thermal treating of the Cd
and/or promoted separation of photoexcited electrons and precursor or post-dipping in a QD solution. Heterostructures
holes. The most famous heterostructure is Degussa P25 TiO2 with well-developed facets can also been obtained by in situ
with 80% anatase and 20% rutile. This material has been crystal growth routes. Through crystallographic-oriented
widely used as a benchmark photocatalyst, though the detailed epitaxial growth of the SnO2 nanorods onto pre-grown
structures of anatase and rutile are still not well recognized. a-Fe2O3 nanospindles and nanocubes, branched SnO2/
Basically, the ultra-high photoreactivity of P25 is originated a-Fe2O3 heterostructures with (101)SnO2/(110)a Fe2O3 and
from the synergistic effects of anatase and rutile mixed (110)SnO2/(110)a Fe2O3 interfaces. respectively, were obtained.188
phases.180–182 Intensive studies have focused on mimicking Due to the different lattice mismatch in the interfaces, SnO2
the synergistic effects by combining anatase and rutile in a nanorods grow at an angle of 90 and 651 (see Fig. 16) with
variety of ways. For example, Tada and co-workers showed respect to exposed facets of nanospindles and nanocubes,
the electron transfer from the conduction band of anatase to respectively. The capability of decomposing methylene blue
the CB of rutile in patterned anatase/rutile bilayer-type junc- with the heterostructures under both UV and visible light is
tion photocatalysts and a resultant much higher photoactivity greatly improved in contrast to sole a-Fe2O3 counterparts. In
in decomposing CH3CHO than with a pure anatase or rutile addition, without any seeds, one-step synthesis of nano-micro
photocatalyst.183 Similarly, Li and co-workers also found that ‘chestnut’ TiO2 with rutile nanopins on the anatase micro-
the junction of anatase nanoparticles on micron-size rutile octahedron was reported by Zhou and co-workers.68
particles can be very effective in promoting photocatalytic
hydrogen evolution from methanol solution.184 The key to 3.8 Introducing dopants in well-faceted photocatalysts
realizing the synergistic effects in heterostructures requires the It is desirable but challenging to design visible-light responsive
intimate contact of the two phases. photocatalysts. Doping, in particular nonmetal doping,49,189–194
Currently, most heterostructures employ hetero-units with- is an important strategy for modifying electronic band struc-
out well-defined facets and thus no clear interfacial atomic ture and thus light response range. The current bottleneck in
structures can be recognized. One apparent drawback is a less
intimate contact among hetero-units. This would affect the

Fig. 16 HRTEM micrographs of SnO2/a-Fe2O3 (nanocube) semi-


conductor nanoheterostructures (SNHs) at different reaction stages:
(a) having reacted for 90 min; (b) having reacted for 120 min, showing
Fig. 15 SEM images of (A) ZnO nanorods and (B) (ZnO)1–(CdS)0.2 slanted growth of SnO2 nanorods on surfaces of a-Fe2O3 nanocubes
core–shell nanorods. Reprinted with permission from ref. 186. with a fixed angle of ca. 651. Reprinted with permission from ref. 188.
Copyright 2010, Elsevier. Copyright 2010, American Chemical Society.

6774 Chem. Commun., 2011, 47, 6763–6783 This journal is c The Royal Society of Chemistry 2011
View Article Online

introducing visible light absorption by doping is that


well-faceted anatase TiO2 crystals usually have very high
crystallinity, making it hard or nearly impossible to incorpo-
rate dopants into their structural framework by mild post-
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

treatment, while the addition of dopant precursors in the


reaction medium may inevitably influence the nucleation and
growth of anatase TiO2 crystals so that no TiO2 sheets with
desirable properties could be synthesized.126
Recently, a new synthesis route has been developed for the
incorporation of nitrogen dopant into anatase crystals with
ca. 60% {001} facets.126 This visible-light photocatalyst has an
absorption band between 400 and 570 nm. The key strategy in
this route is the use of a crystalline compound, titanium nitride
(TiN), as both a precursor and a source of nitrogen dopant.
Fig. 17 Structural models of step edges A–E on anatase (101). For
The method has been extended to prepare sulfur doped
steps A–C, two possible terminations are considered. Some specific Ti
anatase crystals with preferential {001} facets by employing and O atoms at the edges are indicated in the insets. Reprinted with
a TiS2 precursor.133 permission from ref. 59. Copyright 2006, Nature Publishing Group.
Self-doping such as oxygen vacancy and Ti3+ in TiO2 can
not only enhance absorption of visible-light but also promote reactivity of a step is very similar to that of an extended
the separation of photoexcited electrons and holes. Self- surface.59 Therefore, the characteristics of steps on them
doping can be usually imparted by post-treatment (i.e. plas- substantially affect their surface chemistry besides the intrinsic
mon treatment).195 For example, oxygen deficient anatase atomic structures of crystal facets, which is very indicative for
crystals with ca. 56% {001}, whose average particle size is 4 understanding reactivity of photocatalysts with tailored facets.
mm, can be prepared by one-pot hydrolysis of crystalline It is generally considered that nonpolar and autocompen-
metallic TiB2 in HF solution.128 The central feature for sated metal oxide surfaces should be stable with only limited
in situ introducing oxygen deficiency is that gaseous hydrogen relaxations of the ions in the surface region and there
released from the hydrolysis of TiB2 can act as powerful would be no reconstruction.197 However, the stable (1  4)
reductive agent for partially reducing anatase crystals reconstruction in experimental conditions has been discovered
generated. This strategy may also be used for preparing other on the nonpolar and autocompensated anatase (001)
oxygen deficient faceted metal oxides. surface.198,199 Subsequent investigation revealed that rows of
Although oxygen vacancies might be reactive sites, the TiO3 species periodically replace rows of surface bridging
removal behavior of oxygen atoms from different crystal facets oxygen of the (1  1) surface (see Fig. 18), namely the
is quite different due to different atomic arrangement. One ‘ad-molecule’ (ADM) model, to release the large surface
typical instance is that twofold coordinated oxygen atoms is tensile stress present on the surface.197 As a result of the
much easier to be removed from rutile (110) than anatase reconstruction, (001) surface is strongly stabilized by reducing
(101).196 The reason for this difference is that the oxygen the surface energy from 0.90 to 0.51 J m 2, but at the same
atoms in the former are bound to fivefold coordinated time the percentage of fivefold coordinated Ti active sites for
titanium atoms while they are connected to sixfold coordi- water dissociation is decreased by 25%. In addition, (1  2)
nated titanium atoms in the latter. and (1  3) reconstructions were studied for the anatase (010)
and (103) surfaces.110,200 Obviously, the reconstruction will be
3.9 Surface steps and reconstructions of crystal facets another inevitable factor in determining the real reactivity of
desired facets.
Surface steps and reconstructions always exist in crystals. The The above results are related to cases of surface defects on
presence of steps and reconstructions greatly affects surface single crystal surfaces with different orientations. By employing
structure and thus the surface chemistry of crystal facets. The Raman spectroscopy, surface structures of photocatalysts with
microcosmic structures of various metal oxides in particular defined facets can be elucidated to some extent. As listed in
TiO2 have long been investigated in depth.58 Table 1, fluorine species have been widely used to stabilize
Compared to surface point defects, steps and reconstructions
are more complex due to their two-dimensional characteristics.
Edges of monoatomic-height steps at terraces on crystal facets
can be virtually considered as very narrow slices of surfaces.
Furthermore, both the stability and reactivity of a step edge
follows that of the corresponding flat surface as demonstrated
by Gong et al.59 By combining the first-principles calculations
with experiments, it was discovered that (100)-, (112)-, (110)- Fig. 18 (a) Relaxed (001)-(1  1) surface of TiO2 anatase. (b) Relaxed
and (001)-like step edges may exist on anatase TiO2 (101) structure of the ‘ad-molecule’ (ADM) (1  4) reconstruction. (c)
(see Fig. 17), though the later two would be less stable as a Projection of the atomic positions of the ADM model on the plane
result of higher edge surface energies. Further calculations of perpendicular to the ŷ direction. Reprinted with permission from
water adsorbed on different step structures suggest that the ref. 197. Copyright 2001, American Physical Society.

This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 6763–6783 6775
View Article Online

anatase TiO2 {001} facets by forming surface Ti–F bonds.


These surface Ti–F bonds will generate twofold effects:129 (i)
changing fivefold coordinated Ti atoms to sixfold coordinated
Ti atoms; (ii) modifications to the force constants for the
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

surface Ti–O–Ti network imposed by the presence of the


surface terminating Ti–F bonds. These can be indicated by
the apparently weakened B1g mode (See Fig. 19a) besides the
shifting of the Eg mode at 144 cm 1 towards higher frequency.
As a consequence, the photoreactivity is greatly impaired in
terms of both photooxidation and photoreduction reactions.
Although fluorine-free surfaces can be obtained by simply
calcining samples in air or washing samples in NaOH solution,
one should keep in mind a possibility that the surface structure
of {001} facets has been modified by the removal of fluorine
from the surface, and whether the photoreactivity evidenced
represents the real level of targeted {001}. As indicated from Fig. 20 Schematic diagram of the morphology evolution of rutile
the knowledge accumulated from surface science, different TiO2 nanotubes and their corresponding SEM micrographs.201
surface defects must also exist in the synthetic photocatalysts
with desired facets. Both the type and amount of surface parent solid rutile rods with rectangular cross-sections are
defects in microcosmic scale on facets varies with the synthesis exposed to a corrosive solution under hydrothermal conditions.
route employed so that faceted photocatalysts, even with a The property of anisotropic corrosion of facets is expected
very similar morphology, may show quite different photo- to have interesting applications in developing other hollow
reactivity. An extreme example is that serious surface structure materials with faceted surfaces. The stability of photocatalysts
modifications to anatase TiO2 with faceted {001} and {101} is always important for a long-term retention of photocatalytic
imposed by oxygen deficiency, which is indicated by an activity. Fortunately, most photocatalysis processes proceed
additional Raman mode at 155 cm 1, can lead to an apparently at room temperature in non-corrosive solutions. The
improved photocatalytic hydrogen evolution.128 anisotropic corrosion of facets should not be a major concern
under such mild conditions.
3.10 Unique properties of shaped crystals
3.10.2 Interaction dependence of adsorbates on crystal
3.10.1 Anisotropic corrosion of different facets. The high-
facets. Reactant molecules must be preadsorbed on the surface
energy {001} facets of anatase TiO2 are more prone to
of photocatalysts for photocatalytic reactions to occur. Since
dissolution than {101} facets during hydrothermal reactions
the surface atomic structures of facets are different, the
in a corrosive solution containing HF and H2O2.122 This
adsorption energy and states of adsorbates vary with different
property can be used to prepare hollow structures with large
facets. Water molecules play an important role in photocata-
surface areas. Liu et al.201 reported the preparation of hollow
lytic reactions. Intensive studies on the adsorption behavior of
rutile TiO2 tubes with rectangular cross-sections from solid
water molecules on TiO2 surfaces with different orientations
rutile rods by anisotropic corrosion of (001) along the [001]
were conducted.72,202–204 Very recently, an overall review on
direction in hydrochloric acid (see Fig. 20). The mechanism
the theoretical studies of interaction of water and titania has
mainly involves a much faster dissolving rate of the high-
been given by Sun et al.63 The principal concern is that water
energy {001} ends than the low-energy lateral {110} when the
molecules on titania surface are adsorbed molecularly or
dissociatively, which is critically determined by the coordina-
tion numbers and mutual distances of surface atoms. As far as
the adsorptions of water molecules on anatase (101), (001) and
(010) surfaces are concerned, very different trends were
observed: water can only be molecularly adsorbed on Ti5c
atoms on defect free (101),72 while both (001) and (010) with
100% Ti5c atoms can dissociatively adsorb water molecules at
low coverages.72,204 However, the adsorption behavior of
water can be effectively changed by surface defects. For instance,
the bridging oxygen vacancy on anatase (101) can cause the
dissociation of an isolated water molecule at the Ti4c centre near
the vacancy.205 The dependence of other adsorbates such as
maleic anhydride on facets has also been investigated.206
Taking advantage of the different adsorption behaviors of
molecules and ions on crystal facets, it is possible to selectively
Fig. 19 Raman spectra between 300–800 cm 1 of (a) as-prepared, deposit metal nanoparticles on different facets of metal
(b) calcined micron-sized anatase TiO2 with preferential {001} facets, oxides. Recently, Read et al.207 used this approach to
and (c) reference anatase TiO2 without preferential {001} facets.129 construct anisotropic metal oxide–metal interfaces. The atomic

6776 Chem. Commun., 2011, 47, 6763–6783 This journal is c The Royal Society of Chemistry 2011
View Article Online

facet-selective deposition of gold nanoparticles on {100} of fluorine termination can reverse the selective decomposition
Cu2O crystals was achieved by preferential pre-adsorption of order between MB and MO as shown in Fig. 22a. The reversed
sodium dodecyl sulfate (SDS) on {111} as shown in Fig. 21. selectivity is obviously related to the changed surface atomic
Furthermore, it was also shown that when Au(CH3COO)3 was structure by the Ti–F bond, as indicated by the Raman spectra
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

used as the precursor, Au particles were formed on both in Section 3.9. This would affect the adsorption behaviors of
{0001} and {0110}, while Au particles were only found on molecules.
{0110} when AuCl3 was the precursor. The proposed reason
for this difference is the preferential adsorption of Cl on 3.10.4 Reduction and oxidation sites on crystal facets. It
{0001}, which prevents the subsequent deposition of Au. may sound strange to propose that certain crystal facets favor
reduction while others prefer oxidation. However, it is true
3.10.3 Reaction selectivity on different facets. Photocatalytic that the photoexcited electrons and holes can be in principle
processes can degrade organic molecules via the formation of trapped by surface cation and anion sites, i.e. surface Ti4+ and
reactive radicals. However, it is often difficult to selectively O2 ions for TiO2.
transform organic compounds, for example from –CH2OH to It has been demonstrated that Pt particles from the photo-
–CHO.76–79,86 The use of photocatalysis in organic synthesis is reduction of Pt6+ can be deposited only on the {110} facets of
still a challenge. A number of innovative strategies have been rutile TiO2 while PbO2 particles from the photooxidation of
developed to enhance the reaction selectivity of photo- Pd2+ are found mostly on the {011} facets (see Fig. 23a and c).
catalysts.87 These strategies involve either modifying the This suggests that the {110} and {011} facets act as efficient
surface structure (controlling surface electronic charge of reduction and oxidation sites respectively.208 A similar trend
photocatalysts by adjusting pH;87 anchoring surface with was also observed in anatase facets as shown in Fig. 23b and d,
specific molecules for selective adsorption of reactants;88 and though not as clearly as that in rutile.208 The proposed reason
coating a thin layer of molecularly imprinted polymer with for different facets as oxidation and reduction sites is that
recognition ability toward the template molecules)89 or intro- surface energy levels (namely, surface electronic structures) of
ducing a second phase (encapsulating photocatalysts by the conduction bands and valence bands of facets are different
porous silica shell with tunable pores for the diffusion of so that photoexcited electrons and holes can be driven to
reactant molecules;90 and loading photocatalysts on some different facets, leading to the separation of electrons
substrates).80 All these strategies are, however, driven by extrinsic and holes.
exertions instead of intrinsic characters of the photocatalysts. Intrigued by the above results, Zheng et al.177 proposed a
It is also possible to improve the selectivity of a photo- possible hole transfer from {110} and {100} to {111}, together
catalyst by optimizing its intrinsic properties. For example, the with an electron transfer reversely from {111} to {110} and
properties of surface structures inherently determine the {100} in Cu2O microcrystals (see Fig. 24). Moreover, Yanina
adsorption/desorption behaviors of reactant molecules and et al.209 found interesting interfacial electron transfer reactions
reaction intermediates and thus affect significantly the photo- at one surface coupled with those at another via current flow
catalytic selectivity. Morphological control to form different through the crystal bulk in a faceted hematite (a-Fe2O3) single
crystal facets with different surface atomic structures provides crystal. The crystal was exposed to an aqueous solution
a means to tune the selectivity of photocatalysts. Xamena et al.78 containing Fe2+ and oxalate, as shown in Fig. 25. Specifically,
demonstrated the enhanced activity in the shape-selective the edge (hk0) surfaces act directly as an oxidant for Fe2+
photocatalytic degradation of large aromatic molecules with species at the basal (001) surfaces by involving electron
ETS-10 titanosilicate by controlled defect production. Very transfer through the bulk from (001) to (hk0). The driving
recently, Liu et al.135 reported that anatase polyhedra force is provided by the potential difference existing between
with exposed 20% {001} facets show the preferential decom- (001) and (hk0). As a result of this process, (hk0) is gradually
position of methylene blue (MB) with respect to methylene
orange (MO) (see Fig. 22b and c), while P25 titania with below
5% {001} facets favors the decomposition of MO, indicating
the importance of intrinsic surface atomic structure in tuning
photocatalytic selectivity. Such photocatalytic selectivity of
MB (MO) is attributed to the adsorption selectivity towards
MB (MO) molecules. Interestingly, even a partial surface

Fig. 22 A comparison of photocatalytic decomposition of methyl


orange (MO) and methylene blue (MB) by hollow TiO2 microspheres
(HTS) before and after surface modification: (a) as-prepared
fluorinated HTS sample; (b) HTS modified by NaOH washing;
Fig. 21 Utilization of preferential adsorption of additives for shape (c) sample modified by calcination at 600 1C. The catalytic selectivity
control (left) and for atomic plane-selective metal deposition (right). is defined as the ratio (r) of the apparent rate constants (k). Reprinted
Reprinted with permission from ref. 207. Copyright 2009, American with permission from ref. 135. Copyright 2010, American Chemical
Chemical Society. Society.

This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 6763–6783 6777
View Article Online

dissolved and the (001) surface is simultaneously subjected to


an epitaxial growth. However, such a mechanism of electron
transfer between two faces involving the bulk as diffusion path
is established under unique conditions: no light, no oxygen,
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

appropriate growth temperature, presence of Fe2+ in solution,


and a facile dissolution of Fe2O3. Furthermore, the initial
electrons on the (001) face, which will diffuse through the
bulk to (hk0), are released from Fe2+ in aqueous solution.
Regarding photocatalysts, the photoexcited charge carriers,
most of which are generated within bulk, will diffuse from the
bulk towards the surface before recombination. Once reaching
the surface, the carriers are trapped by the topmost exposed
atoms on different facets. Then, the concern is whether or not
the trapped charge carriers on different facets will transfer
from one facet to others as a result of different surface
electronic structure.
The proposed direct charge transfer process from one facet
to others is debatable when one considers the following
factors: (1) two adjacent facets ideally share only one atomic
Fig. 23 SEM images of a rutile particle (a) and an anatase particle line or chain so the probability for transfer between facets
(b) on which Pt fine particles were deposited by UV-irradiation in a should be very small; and (2) the electronic states related to the
solution of 1.0 mM H2PtCl6; SEM images of a rutile particle (c) and an surface terminated atoms are highly localized so the mobility
anatase particle (d) showing PbO2 deposits, which were loaded on the of the trapped charge carriers should be quite low. Therefore,
particles by UV-irradiation of the Pt-deposited TiO2 powder in a most of the trapped charge carriers on different facets should
solution of 0.1 M Pb(NO3)2.208 stay reasonably intact.
Finally but importantly, the possible influence of probing
agents used to reveal the reduction/oxidation facets is often
overlooked. Ionic species such as Cl from H2PtCl6 and NO3
from Pb(NO3)2 may be preferentially adsorbed on some facets,
which can subsequently result in selective deposition of metal
particles on other facets. This possibility has already been
clearly demonstrated in ZnO crystals.175 Furthermore, the
initial solid cluster products from reduced/oxidized ions
deposited on photocatalyst surfaces would accelerate the
subsequent deposition of reduced/oxidized ions on these
clusters. It is uncertain if respective facets for reduction and
Fig. 24 Schematic diagram showing charge separation among
oxidation sites can still exist when molecules such as water,
different crystal faces and regeneration of Cu2O nanosheets. Reprinted
with permission from ref. 177. Copyright 2010, American Chemical phenol and azo dyes are employed as probing agents. The
Society. products are either intermediate species desorbed from the
photocatalyst surface, or mineralized compounds such as CO2,
which have negligible effects on subsequent photocatalytic
reactions.
3.11 Photoreactivity order of different facets and its origin
Determining the photocatalytic reactivity order of different
facets is necessary for designing efficient photocatalysts with
well-defined facets. The conventional understanding of the
surface atomic structure of a crystal is that facets with a higher
percentage of under-coordinated atoms are usually more
reactive in heterogeneous reactions. This sounds reasonable
when one considers that the under-coordinated atoms may act
as active reaction sites. However, the substantial role of the
electronic structures in affecting the photocatalytic reactivity
of facets is often neglected. Due to the difference in surface
atomic arrangements, the band gap of crystal facets or crystal
Fig. 25 Schematic diagram depicting the inferred coupled interfacial slabs would change. This will correspondingly change the
electron transfer process operative under the conditions for the redox power of the photoexcited electrons and holes. As
hematite single crystals. Reprinted with permission from ref. 209. demonstrated by Liu et al.,129 the UV-visible absorption
Copyright 2008, AAAS. spectrum of anatase TiO2 crystals with 82% {101} shows an

6778 Chem. Commun., 2011, 47, 6763–6783 This journal is c The Royal Society of Chemistry 2011
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM. View Article Online

Fig. 26 UV-visible absorption spectra of nanosized anatase crystals


with 18% {001} facets and micron-sized anatase crystals with 72%
{001} facets.129

obvious blue shift compared to that with 72% {001} (see


Fig. 27 (a) Fluorescence signal intensity of TAOH at 426 nm and
Fig. 26), indicating that {101} has a larger bandgap than
(b) hydrogen evolution rate from water containing 10% methanol in
{001}. Further electronic structure analysis suggests that volume for different samples: surface fluorine-terminated anatase TiO2
{101} has the same valence band (VB) maximum as {001}, crystals T001-F, T101-F and T010-F; clean anatase TiO2 crystals
but has a higher conduction band (CB) minimum than {001}. T001, T101 and T010 (T001, T101 and T010 indicates the sample
This electronic band difference, together with a difference in with a respective predominance of {001}, {101} and {010}). Reprinted
atomic coordination, would certainly affect photocatalytic with permission from ref. 75. Copyright 2011, Wiley-VCH.
reactivity.
Very recently, Pan et al. systematically compared photo-
catalytic reactivity in generating  OH radicals and in H2
evolution of anatase crystals enclosed with dominant {001},
{101} and {010}. It was found that despite of the different
surface atomic structure, all three facets had similar photo-
catalytic reactivity when partially terminated with fluorine.
Contrary to conventional understanding, {001} exhibited a
lower reactivity than {101}, and {010} demonstrated the high-
est photoreactivity (see Fig. 27). For the surface under-
coordinated Ti5c atoms, both {001} and {010} exhibited a
higher photoreactivity than {101}. Apparently, {010} facets
possess both a favorable surface atomic structure and a
surface electronic structure so that the stronger reducing Fig. 28 (A) Structure of anatase TiO2 crystal with preferential (001)
electrons on CB can be transferred via the surface Ti5c atoms facets. Transmission image (B) of a single TiO2 crystal on the
as active reaction sites. Such a cooperative mechanism existing cover glass in Ar-saturated 2.0 mm 3,4-dinitrophenyl-BODIPY
on {010} facets is responsible for its unexpectedly high reactivity. (DN-BODIPY) solution under 488 nm laser and UV irradiation.
The higher photocatalytic reactivity of {101} than {001} has The red and blue dots in image (B) indicate the fluorescence bursts
also been validated by probing single molecule fluorescence located on the (101) and (001) surfaces, respectively, observed during 3
min irradiation. Reprinted with permission from ref. 210. Copyright
burst on a single anatase crystal particle.210,211 As shown in
2010, Wiley-VCH.
Fig. 28 by Tachikawa et al.,210 the red and blue dots represent
the fluorescence bursts on (101) and (001), respectively. The
number of red dots on {101} is about three times higher than As presented in Section 3.9, different defects always occur on
the number of blue dots on (001), although the surface area of different facets. Moreover, the abundance and nature of
(001) is more than two times higher than that of (101).129 defects are related to the synthesis routes employed. It is
Based on the above results and discussion, a high density of therefore difficult to compare the activity of photocatalysts
surface under-coordinated atoms does cause a high surface obtained from different synthesis routes.
energy for the crystal facet, but it does not always make the
crystal facet highly reactive in photocatalytic reactions.
4. Conclusions
A crystal facet with a high density of under-coordinated atoms
can be photocatalytically advantageous only if a favourable Crystal facet engineering has attracted increasing attention in
electronic structure co-exists. photocatalysis since Yang et al. reported an anatase TiO2 with
Finally, the possible influence of surface defects on the a large percentage of high surface-energy {001} facets.21 Most
photocatalytic reactivity of facets should also be considered. of the studies focused on developing various synthesis routes

This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 6763–6783 6779
View Article Online

to prepare pure anatase crystals with higher percentage of electrons and holes. Doping and heterostructuring are strategies
{001} and smaller particle size. Anion doping in anatase that will definitely contribute to the development of more
crystals has also been explored. The wet-chemistry methods efficient faceted photocatalysts. Doping changes the surface
are most popular but many solid-state and chemical vapor atomic structure of facets and thus alters their reactivity.
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

methods have also been developed. As a result, the percentage Heterostructuring allows effective interfacial charge carrier
of {001} can now approach 100% and the thickness of anatase transfer by constructing favorable interfaces with various
crystals is as thin as 1.6 nm. Meanwhile, the interest has facets from different crystals. The key points related to
already spread from TiO2 to other metal oxide photocatalysts. growing, doping and heterostructuring of faceted photo-
These faceted photocatalysts have already been used in photo- catalysts are summarized in Fig. 29.
degradation of organic pollutants, photocatalytic water
splitting and even dye-sensitized solar cells. Facet related Acknowledgements
photocatalytic reactivity and selectivity improvements will
have implications on organic synthesis. Unique properties The work described in this paper was partially supported
including anisotropic facet corrosion and preferential by the NSFC-RGC Joint Research Scheme (Project
molecule adsorption have been used to produce hollow N_CUHK464/10), the Major Basic Research Program,
structures and anisotropic heterostructures. Importantly, the Ministry of Science and Technology of China (No.
understanding of high reactive facets has been extended from 2009CB220001), NSFC (Nos. 50921004, 51002160 and
the conventional view of a surface with a high density of 21090343), Solar Energy Initiative and the Funding
under-coordinated atoms to the cooperative mechanism of (KJCX2-YW-H21-01) of the Chinese Academy of Sciences
both favourable surface atomic structure and surface for financial support. GL thanks the IMR SYNL-T.S. Kê
electronic structure. In addition, new technology such as single Research Fellowship.
molecule fluorescence probing has facilitated the measurement
of photocatalytic activity for each facet. Notes and references
Although encouraging advances in synthesizing photo-
1 S. E. Habas, H. Lee, V. Radmilovic, G. A. Somorjai and P. Yang,
catalysts with defined facets has been achieved in recent years, Nat. Mater., 2007, 6, 692.
most effort has focused on binary photocatalysts, particularly 2 C. K. Tsung, J. N. Kuhn, W. Y. Huang, C. Aliaga, L. I. Hung,
TiO2. Faceted ternary photocatalysts are still challenging yet G. A. Somorjai and P. D. Yang, J. Am. Chem. Soc., 2009, 131,
desirable to be explored for crystal facet engineering to fully 5816.
3 A. Tao, P. Sinsermsuksakul and P. D. Yang, Angew. Chem., Int.
exert its influence. Continuous efforts are needed to develop Ed., 2006, 45, 4597.
feasible synthesis routes for growing crystals with targeted 4 A. I. Hochbaum and P. D. Yang, Chem. Rev., 2010, 110, 527.
reactive facets. Most synthesis strategies involve the use of 5 Y. Xia, Y. J. Xiong, B. Lim and S. E. Skrabalak, Angew. Chem.,
Int. Ed., 2008, 48, 60.
morphology controlling agents (MCAs) that must be 6 B. Wiley, Y. G. Sun, J. Y. Chen, H. Cang, Z. Y. Li, X. D. Li and
eventually removed in order to obtain clean facets. This Y. N. Xia, MRS Bull., 2011, 30, 356.
process may cause some uncontrollable changes to surface 7 B. Lim, H. Kobayashi, T. Yu, J. G. Wang, M. J. Kim, Z. Y. Li,
atomic structures. The development of MCA-free synthesis M. Rycenga and Y. Xia, J. Am. Chem. Soc., 2010, 132, 2506.
8 Y. J. Xiong and Y. N. Xia, Adv. Mater., 2007, 19, 3385.
routes is therefore highly desirable. 9 B. Wiley, Y. G. Sun, B. Mayers and Y. N. Xia, Chem.–Eur. J.,
Many stable photocatalysts suffer from weak visible-light 2005, 11, 454.
absorption and high recombination rate of photoexcited 10 B. Sadtler, D. O. Demchenko, H. Zheng, S. M. Hughes,
M. G. Merkle, U. Dahmen, L. W. Wang and A. P. Alivisatos,
J. Am. Chem. Soc., 2009, 131, 5285.
11 Y. D. Yin, C. Erdonmez, S. Aloni and A. P. Alivisatos, J. Am.
Chem. Soc., 2006, 128, 12671.
12 X. J. Feng, J. Zhai and L. Jiang, Angew. Chem., Int. Ed., 2005, 44,
5115.
13 X. L. Li, Q. Peng, J. X. Yi, X. Wang and Y. D. Li, Chem.–Eur. J.,
2006, 12, 2383.
14 X. Wang, J. Zhuang, Q. Peng and Y. D. Li, Nature, 2005, 437,
121.
15 Y. G. Sun and Y. N. Xia, Science, 2002, 298, 2176.
16 F. Wang, Y. Han, C. S. Lim, Y. H. Lu, J. Wang, J. Xu,
H. Y. Chen, C. Zhang, M. H. Hong and X. G. Liu, Nature,
2010, 463, 1061.
17 B. Lim, M. J. Jiang, P. H. C. Camargo, E. C. Cho, J. Tao,
X. M. Lu, Y. M. Zhu and Y. A. Xia, Science, 2009, 324, 1302.
18 N. Tian, Z. Y. Zhou, S. G. Sun, Y. Ding and Z. L. Wang, Science,
2007, 316, 732.
19 X. W. Xie, Y. Li, Z. Q. Liu, M. Haruta and W. J. Shen, Nature,
2009, 458, 746.
20 X. D. Feng, D. C. Sayle, Z. L. Wang, M. S. Paras, B. Santora,
A. C. Sutorik, T. X. T. Sayle, Y. Yang, Y. Ding, X. D. Wang and
Y. S. Her, Science, 2006, 312, 1504.
21 H. G. Yang, C. H. Sun, S. Z. Qiao, J. Zou, G. Liu, S. C. Smith,
H. M. Cheng and G. Q. Lu, Nature, 2008, 453, 638.
Fig. 29 Highlighted key points in growing, doping and hetero- 22 X. G. Peng, L. Manna, W. D. Yang, J. Wickham, E. Scher,
structuring of faceted semiconductor photocatalysts. A. Kadavanich and A. P. Alivisatos, Nature, 2000, 404, 59.

6780 Chem. Commun., 2011, 47, 6763–6783 This journal is c The Royal Society of Chemistry 2011
View Article Online

23 J. H. Xiang, S. H. Yu, B. H. Liu, Y. Xu, X. Gen and L. Ren, 63 C. H. Sun, L. M. Liu, A. Selloni, G. Q. Lu and S. C. Smith,
Inorg. Chem. Commun., 2004, 7, 572. J. Mater. Chem., 2010, 20, 10319.
24 C. Z. Wu and Y. Xie, Chem. Commun., 2009, 5943. 64 L. Kavan, M. Gratzel, S. E. Gilbert, C. Klemenz and H. J. Scheel,
25 X. G. Han, M. S. Jin, S. F. Xie, Q. Kuang, Z. Y. Jiang, J. Am. Chem. Soc., 1996, 118, 6716.
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

Y. Q. Jiang, Z. X. Xie and L. S. Zheng, Angew. Chem., Int. 65 E. S. Jang, J. H. Won, S. J. Hwang and J. H. Choy, Adv. Mater.,
Ed., 2009, 48, 9180. 2006, 18, 3309.
26 C. Burda, X. B. Chen, R. Narayanan and M. A. El-Sayed, Chem. 66 R. L. Penn and J. F. Banfield, Geochim. Cosmochim. Acta, 1999,
Rev., 2005, 105, 1025. 63, 1549.
27 X. Chen and S. S. Mao, Chem. Rev., 2007, 107, 2891. 67 A. Chemseddine and T. Moritz, Eur. J. Inorg. Chem., 1999, 235.
28 H. C. Zeng, J. Mater. Chem., 2006, 16, 649. 68 E. Hosono, S. Fujihara, H. Lmai, I. Honma, I. Masaki and
29 H. G. Yang and H. C. Zeng, Angew. Chem., Int. Ed., 2004, 43, 5930. H. S. Zhou, ACS Nano, 2007, 1, 273.
30 A. Fujishima and K. Honda, Nature, 1972, 238, 37. 69 C. T. Dinh, T. D. Nguyen, F. Kleitz and T. O. Do, ACS Nano,
31 M. R. Hoffmann, S. T. Martin, W. Y. Choi and D. W. 2009, 3, 3737.
Bahnemann, Chem. Rev., 1995, 95, 69. 70 Z. H. Zhang, X. H. Zhong, S. H. Liu, D. F. Li and M. Y. Han,
32 A. L. Linsebigler, G. Q. Lu and J. T. Yates, Chem. Rev., 1995, 95, Angew. Chem., Int. Ed., 2005, 44, 3466.
735. 71 X. Q. Gong and A. Selloni, J. Phys. Chem. B, 2005, 109, 19560.
33 X. B. Chen, S. H. Shen, L. J. Guo and S. S. Mao, Chem. Rev., 72 A. Vittadini, A. Selloni, F. P. Rotzinger and M. Gratzel, Phys.
2010, 110, 6503. Rev. Lett., 1998, 81, 2954.
34 A. Kudo and Y. Miseki, Chem. Soc. Rev., 2009, 38, 253. 73 A. Selloni, Nat. Mater., 2008, 7, 613.
35 H. Tada, T. Kiyonaga and S. Naya, Chem. Soc. Rev., 2009, 38, 74 A. Vittadini, M. Casarin and A. Selloni, Theor. Chem. Acc., 2006,
1849. 117, 663.
36 K. Maeda and K. Domen, Chem. Mater., 2010, 22, 612. 75 J. Pan, G. Liu, G. Q. Lu and H. M. Cheng, Angew. Chem., Int.
37 G. Liu, L. Z. Wang, H. G. Yang, H. M. Cheng and G. Q. Lu, Ed., 2011, 50, 2133.
J. Mater. Chem., 2010, 20, 831. 76 X. T. Shen, L. H. Zhu, G. X. Liu, H. W. Yu and H. Q. Tang,
38 Z. G. Zou, J. H. Ye, K. Sayama and H. Arakawa, Nature, 2001, Environ. Sci. Technol., 2008, 42, 1687.
414, 625. 77 S. Yurdakal, G. Palmisano, V. Loddo, V. Augugliaro and
39 K. Maeda, K. Teramura, D. L. Lu, T. Takata, N. Saito, Y. Inoue L. Palmisano, J. Am. Chem. Soc., 2008, 130, 1568.
and K. Domen, Nature, 2006, 440, 295. 78 F. X. L. I. Xamena, P. Calza, C. Lamberti, C. Prestipino,
40 T. Inoue, A. Fujishima, S. Konishi and K. Honda, Nature, 1979, A. Damin, S. Bordiga, E. Pelizzetti and A. Zecchina, J. Am.
277, 637. Chem. Soc., 2003, 125, 2264.
41 B. Oregan and M. Gratzel, Nature, 1991, 353, 737. 79 Y. Shiraishi and T. Hirai, J. Photochem. Photobiol., C, 2008, 9,
42 M. Gratzel, Nature, 2001, 414, 338. 157.
43 H. J. Zhang, G. H. Chen and D. W. Bahnemann, J. Mater. 80 H. Uchida, S. Itoh and H. Yoneyama, Chem. Lett., 1993, 1995.
Chem., 2009, 19, 5089. 81 X. H. Xia, Z. H. Jia, Y. Yu, Y. Liang, Z. Wang and L. L. Ma,
44 Z. G. Yi, J. H. Ye, N. Kikugawa, T. Kako, S. X. Ouyang, Carbon, 2007, 45, 717.
H. Stuart-Williams, H. Yang, J. Y. Cao, W. J. Luo, Z. S. Li, 82 G. R. Dey, J. Nat. Gas Chem., 2007, 16, 217.
Y. Liu and R. L. Withers, Nat. Mater., 2010, 9, 559. 83 V. P. Indrakanti, J. D. Kubicki and H. H. Schobert, Energy
45 D. Q. Zhang, G. S. Li and J. C. Yu, J. Mater. Chem., 2010, 20, Environ. Sci., 2009, 2, 745.
4529. 84 S. C. Roy, O. K. Varghese, M. Paulose and C. A. Grimes, ACS
46 H. J. Yan, J. H. Yang, G. J. Ma, G. P. Wu, X. Zong, Z. B. Lei, Nano, 2010, 4, 1259.
J. Y. Shi and C. Li, J. Catal., 2009, 266, 165. 85 P. Calza, C. Paze, E. Pelizzetti and A. Zecchina, Chem. Commun.,
47 X. Z. Fu, L. A. Clark, W. A. Zeltner and M. A. Anderson, 2001, 2130.
J. Photochem. Photobiol., A, 1996, 97, 181. 86 S. Usseglio, P. Calza, A. Damin, C. Minero, S. Bordiga,
48 X. C. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, C. Lamberti, E. Pelizzetti and A. Zecchina, Chem. Mater., 2006,
J. M. Carlsson, K. Domen and M. Antonietti, Nat. Mater., 18, 3412.
2008, 8, 76. 87 D. Robert, A. Piscopo and J. V. Weber, Sol. Energy, 2004, 77,
49 W. Zhao, W. H. Ma, C. C. Chen, J. C. Zhao and Z. G. Shuai, 553.
J. Am. Chem. Soc., 2004, 126, 4782. 88 O. V. Makarova, T. Rajh, M. C. Thurnauer, A. Martin, P. A.
50 J. C. Yu, J. G. Yu, W. K. Ho, Z. T. Jiang and L. Z. Zhang, Chem. Kemme and D. Cropek, Environ. Sci. Technol., 2000, 34, 4797.
Mater., 2002, 14, 3808. 89 X. T. Shen, L. H. Zhu, J. Li and H. Q. Tang, Chem. Commun.,
51 O. K. Varghese, M. Paulose, T. J. LaTempa and C. A. Grimes, 2007, 1163.
Nano Lett., 2009, 9, 731. 90 S. Wang, T. Wang, W. X. Chen and T. R. Hori, Chem. Commun.,
52 G. K. Mor, H. E. Prakasam, O. K. Varghese, K. Shankar and 2008, 3756.
C. A. Grimes, Nano Lett., 2007, 7, 2356. 91 M. A. Lovette, A. R. Browning, D. W. Griffin, J. P. Sizemore,
53 H. G. Kim, P. H. Borse, W. Y. Choi and J. S. Lee, Angew. Chem., R. C. Snyder and M. F. Doherty, Ind. Eng. Chem. Res., 2008, 47,
Int. Ed., 2005, 44, 4585. 9812.
54 Y. X. Li, G. X. Lu and S. B. Li, Appl. Catal., A, 2001, 214, 179. 92 A. Seyed-Razavi, I. K. Snook and A. S. Barnard, J. Mater.
55 L. S. Zhang, W. Z. Wang, L. Zhou and H. L. Xu, Small, 2007, 3, Chem., 2010, 20, 416.
1618. 93 J. Y. Chen, B. Lim, E. P. Lee and Y. N. Xia, Nano Today, 2009,
56 X. J. Lu, F. Q. Huang, X. L. Mou, Y. M. Wang and F. F. Xu, 4, 81.
Adv. Mater., 2010, 22, 3719. 94 K. Lee, M. Kim and H. Kim, J. Mater. Chem., 2010, 20, 3791.
57 Q. Wang, M. A. Zhang, C. C. Chen, W. H. Ma and J. C. Zhao, 95 Z. Y. Jiang, Q. Kuang, Z. X. Xie and L. S. Zheng, Adv. Funct.
Angew. Chem., Int. Ed., 2010, 49, 7976. Mater., 2010, 20, 3634.
58 U. Diebold, Surf. Sci. Rep., 2003, 48, 53. 96 D. Kashchiev, P. G. Vekilov and A. B. Kolomeisky, J. Chem.
59 X. Q. Gong, A. Selloni, M. Batzill and U. Diebold, Nat. Mater., Phys., 2005, 122, 244706.
2006, 5, 665. 97 P. G. Vekilov, Cryst. Growth Des., 2004, 4, 671.
60 Y. B. He, A. Tilocca, O. Dulub, A. Selloni and U. Diebold, Nat. 98 W. C. Pan, A. B. Kolomeisky and P. G. Vekilov, J. Chem. Phys.,
Mater., 2009, 8, 585. 2005, 122, 174905.
61 O. Dulub, M. Batzill, S. Solovev, E. Loginova, A. Alchagirov, 99 T. H. Zhang and X. Y. Liu, J. Am. Chem. Soc., 2007, 129, 13520.
T. E. Madey and U. Diebold, Science, 2007, 317, 1052. 100 R. J. Davey, J. Cryst. Growth, 1986, 76, 637.
62 (a) S. C. Li, L. N. Chu, X. Q. Gong and U. Diebold, Science, 101 J. R. Bourne and R. J. Davey, J. Cryst. Growth, 1976, 36, 278.
2010, 328, 882; (b) C. Y. Zhou, Z. F. Ren, S. J. Tan, Z. B. Ma, 102 J. R. Bourne and R. J. Davey, J. Cryst. Growth, 1976, 36, 287.
X. C. Mao, D. X. Dai, H. J. Fan, X. M. Yang, J. LaRue, 103 M. Lahav and L. Leiserowitz, Chem. Eng. Sci., 2001, 56, 2245.
R. Cooper, A. M. Wodtke, Z. Wang, Z. Y. Li, B. Wang, 104 D. Lechugaballesteros and N. Rodriguezhornedo, Int. J. Pharm.,
J. L. Yang and J. G. Hou, Chem. Sci., 2010, 1, 575. 1995, 115, 139.

This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 6763–6783 6781
View Article Online

105 J. S. Chen, T. Zhu, X. H. Yang, H. G. Yang and X. W. Lou, 141 J. Zhu, S. H. Wang, Z. F. Bian, S. H. Xie, C. L. Cai, J. G. Wang,
J. Am. Chem. Soc., 2010, 132, 13162. H. G. Yang and H. X. Li, CrystEngComm, 2010, 12, 2219.
106 Z. G. Zhao, Z. F. Liu and M. Miyauchi, Chem. Commun., 2010, 142 J. G. Yu, Q. J. Xiang, J. R. Ran and S. Mann, CrystEngComm,
46, 3321. 2010, 12, 872.
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

107 T. Taguchi, Y. Saito, K. Sarukawa, T. Ohno and M. Matsumura, 143 X. W. Wang, G. Liu, L. Z. Wang, J. Pan, G. Q. Lu and
New J. Chem., 2003, 27, 1304. H. M. Cheng, J. Mater. Chem., 2011, 21, 869.
108 Z. Y. Tang, N. A. Kotov and M. Giersig, Science, 2002, 297, 237. 144 L. Zhou, D. Smyth-Boyle and P. O’Brien, J. Am. Chem. Soc.,
109 C. Schliehe, B. H. Juarez, M. Pelletier, S. Jander, D. Greshnykh, 2008, 130, 1309.
M. Nagel, A. Meyer, S. Foerster, A. Kornowski, C. Klinke and 145 B. H. Wu, C. Y. Guo, N. F. Zheng, Z. X. Xie and G. D. Stucky,
H. Weller, Science, 2010, 329, 550. J. Am. Chem. Soc., 2008, 130, 17563.
110 M. Lazzeri, A. Vittadini and A. Selloni, Phys. Rev. B: Condens. 146 J. M. Li and D. S. Xu, Chem. Commun., 2010, 46, 2301.
Matter, 2001, 63, 155409. 147 P. H. Wen, Y. Ishikawa, H. Itoh and Q. Feng, J. Phys. Chem. C,
111 M. Lazzeri, A. Vittadini and A. Selloni, Phys. Rev. B: Condens. 2009, 113, 20275.
Matter, 2002, 65, 119901. 148 J. Zhang, Q. Xu, M. J. Li, Z. C. Feng and C. Li, J. Phys. Chem. C,
112 M. Ramamoorthy, D. Vanderbilt and R. D. Kingsmith, Phys. 2009, 113, 1698.
Rev. B, 1994, 49, 1672. 149 G. Liu, F. Li, Z. G. Chen, G. Q. Lu and H. M. Cheng, J. Solid
113 X. Q. Gong and A. Selloni, Phys. Rev. B: Condens. Matter Mater. State Chem., 2006, 179, 331.
Phys., 2007, 76, 235307. 150 J. M. Wu, H. C. Shih and W. T. Wu, Nanotechnology, 2006, 17,
114 A. S. Barnard, P. Zapol and L. A. Curtiss, J. Chem. Theory 105.
Comput., 2005, 1, 107. 151 C. R. Xiong, X. Y. Deng and J. B. Li, Appl. Catal., B, 2010, 94,
115 A. S. Barnard and L. A. Curtiss, Nano Lett., 2005, 5, 1261. 234.
116 F. Amano, T. Yasumoto, O. O. Prieto-Mahaney, S. Uchida, 152 D. Q. Zhang, G. S. Li, F. Wang and J. C. Yu, CrystEngComm,
T. Shibayama and B. Ohtani, Chem. Commun., 2009, 2311. 2010, 12, 1759.
117 Y. Q. Dai, C. M. Cobley, J. Zeng, Y. M. Sun and Y. N. Xia, Nano 153 K. Kakiuchi, E. Hosono, H. Imai, T. Kimura and S. Fujihara,
Lett., 2009, 9, 2455. J. Cryst. Growth, 2006, 293, 541.
118 G. Liu, H. G. Yang, C. H. Sun, L. N. Cheng, L. Z. Wang, 154 X. F. Yang, J. Chen, L. Gong, M. M. Wu and J. C. Yu, J. Am.
G. Q. Lu and H. M. Cheng, CrystEngComm, 2009, 11, 2677. Chem. Soc., 2009, 131, 12048.
119 N. Q. Wu, J. Wang, D. Tafen, H. Wang, J. G. Zheng, J. P. Lewis, 155 Y. W. Wang, L. Z. Zhang, K. J. Deng, X. Y. Chen and Z. G. Zou,
X. G. Liu, S. S. Leonard and A. Manivannan, J. Am. Chem. Soc., J. Phys. Chem. C, 2007, 111, 2709.
2010, 132, 6679. 156 H. M. Cheng, J. M. Ma, Z. G. Zhao and L. M. Qi, Chem. Mater.,
120 X. G. Han, Q. Kuang, M. S. Jin, Z. X. Xie and L. S. Zheng, 1995, 7, 663.
J. Am. Chem. Soc., 2009, 131, 3152. 157 Y. Sang, B. Y. Geng and J. Yang, Nanoscale, 2010, 2, 2109–2113.
121 H. G. Yang, G. Liu, S. Z. Qiao, C. H. Sun, Y. G. Jin, S. C. Smith, 158 B. Liu and E. S. Aydil, J. Am. Chem. Soc., 2009, 131, 3985.
J. Zou, H. M. Cheng and G. Q. Lu, J. Am. Chem. Soc., 2009, 131, 159 T. Y. Ke, C. W. Peng, C. Y. Lee, H. T. Chiu and H. S. Sheu,
4078. CrystEngComm, 2009, 11, 1691.
122 X. Y. Hu, T. Zhang, Z. Jin, S. Z. Huang, M. Fang, Y. C. Wu and 160 E. Bae and T. Ohno, Appl. Catal. B: Environ., 2009, 91, 634.
L. Zhang, Cryst. Growth Des., 2009, 9, 2324. 161 J. Yu, Y. Chen and A. M. Glushenkov, Cryst. Growth Des., 2009,
123 D. Q. Zhang, G. S. Li, X. F. Yang and J. C. Yu, Chem. Commun., 9, 1240.
2009, 4381. 162 F. Y. Wei, H. L. Zeng, P. Cui, S. C. Peng and T. H. Cheng, Chem.
124 F. Amano, O. O. Prieto-Mahaney, Y. Terada, T. Yasumoto, Eng. J., 2008, 144, 119.
T. Shibayama and B. Ohtani, Chem. Mater., 2009, 21, 2601. 163 H. Kaper, F. Endres, I. Djerdj, M. Antonietti, B. M. Smarsly,
125 Y. Alivov and Z. Y. Fan, J. Phys. Chem. C, 2009, 113, 12954. J. Maier and Y. S. Hu, Small, 2007, 3, 1753.
126 G. Liu, H. G. Yang, X. W. Wang, L. N. Cheng, J. Pan, G. Q. Lu 164 H. Xu, F. L. Jia, Z. H. Ai and L. Z. Zhang, Cryst. Growth Des.,
and H. M. Cheng, J. Am. Chem. Soc., 2009, 131, 12868. 2007, 7, 1216.
127 Z. K. Zheng, B. B. Huang, X. Y. Qin, X. Y. Zhang, Y. Dai, 165 A. Dessombz, D. Chiche, P. Davidson, P. Panine, C. Chaneac
M. H. Jiang, P. Wang and M. H. Whangbo, Chem.–Eur. J., 2009, and J. P. Jolivet, J. Am. Chem. Soc., 2007, 129, 5904.
15, 12576. 166 M. N. Tahir, P. Theato, P. Oberle, G. Melnyk, S. Faiss, U. Kolb,
128 G. Liu, H. G. Yang, X. W. Wang, L. N. Cheng, H. F. Lu, A. Janshoff, M. Stepputat and W. Tremel, Langmuir, 2006, 22,
L. Z. Wang, G. Q. Lu and H. M. Cheng, J. Phys. Chem. C, 2009, 5209.
113, 21784. 167 M. Alfredsson, F. Cora, D. P. Dobson, J. Davy, J. P. Brodholt,
129 G. Liu, C. H. Sun, H. G. Yang, S. C. Smith, L. Z. Wang, G. Q. Lu S. C. Parker and G. D. Price, Surf. Sci., 2007, 601, 4793.
and H. M. Cheng, Chem. Commun., 2010, 46, 755. 168 B. Zhao, F. Chen, Q. W. Huang and J. L. Zhang, Chem.
130 M. Liu, L. Y. Piao, L. Zhao, S. T. Ju, Z. J. Yan, T. He, Commun., 2009, 5115.
C. L. Zhou and W. J. Wang, Chem. Commun., 2010, 46, 1664. 169 M. Addamo, V. Augugliaro, M. Bellardita, A. Di Paola,
131 J. S. Chen, Y. L. Tan, C. M. Li, Y. L. Cheah, D. Y. Luan, V. Loddo, G. Palmisano, L. Palmisano and S. Yurdakal, Catal.
S. Madhavi, F. Y. C. Boey, L. A. Archer and X. W. Lou, J. Am. Lett., 2008, 126, 58.
Chem. Soc., 2010, 132, 6124. 170 F. Iskandar, A. B. D. Nandiyanto, K. M. Yun, C. J. Hogan,
132 J. G. Yu, L. F. Qi and M. Jaroniec, J. Phys. Chem. C, 2010, 114, K. Okuyama and P. Biswas, Adv. Mater., 2007, 19, 1408.
13118. 171 J. G. Li, C. C. Tang, D. Li, H. Haneda and T. Ishigaki, J. Am.
133 G. Liu, C. H. Sun, S. C. Smith, L. Z. Wang, G. Q. Lu and Ceram. Soc., 2004, 87, 1358.
H. M. Cheng, J. Colloid Interface Sci., 2010, 349, 477. 172 R. Buonsanti, V. Grillo, E. Carlino, C. Giannini, T. Kipp,
134 X. Y. Ma, Z. G. Chen, S. B. Hartono, H. B. Jiang, J. Zou, R. Cingolani and P. D. Cozzoli, J. Am. Chem. Soc., 2008, 130, 11223.
S. Z. Qiao and H. G. Yang, Chem. Commun., 2010, 46, 6608. 173 G. C. Xi and J. H. Ye, Chem. Commun., 2010, 46, 1893.
135 S. W. Liu, J. G. Yu and M. Jaroniec, J. Am. Chem. Soc., 2010, 174 J. H. Zeng, B. B. Jin and Y. F. Wang, Chem. Phys. Lett., 2009,
132, 11914. 472, 90.
136 X. G. Liu, D. Y. Geng, X. L. Wang, S. Ma, H. Wang, D. Li, 175 A. Mclaren, T. Valdes-Solis, G. Q. Li and S. C. Tsang, J. Am.
B. Q. Li, W. Liu and Z. D. Zhang, Chem. Commun., 2010, 46, 6956. Chem. Soc., 2009, 131, 12540.
137 J. Y. Feng, M. C. Yin, Z. Q. Wang, S. C. Yan, L. J. Wan, Z. S. Li 176 W. W. Zhou, B. Yan, C. W. Cheng, C. X. Cong, H. L. Hu,
and Z. G. Zou, CrystEngComm, 2010, 12, 3425. H. J. Fan and T. Yu, CrystEngComm, 2009, 11, 2291.
138 J. G. Yu, J. J. Fan and K. L. Lv, Nanoscale, 2010, 2, 2144. 177 Y. Zhang, B. Deng, T. R. Zhang, D. M. Gao and A. W. Xu,
139 H. M. Zhang, Y. H. Han, X. L. Liu, P. R. Liu, H. Yu, J. Phys. Chem. C, 2010, 114, 5073.
S. Q. Zhang, X. D. Yao and H. J. Zhao, Chem. Commun., 178 D. E. Wang, H. F. Jiang, X. Zong, Q. Xu, Y. Ma, G. L. Li and
2010, 46, 8395. C. Li, Chem.–Eur. J., 2011, 17, 1275.
140 D. Q. Zhang, G. S. Li, H. B. Wang, K. M. Chan and J. C. Yu, 179 Q. Liu, Y. Zhou, J. H. Kou, X. Y. Chen, Z. P. Tian, J. Gao,
Cryst. Growth Des., 2010, 10, 1130. S. C. Yan and Z. G. Zou, J. Am. Chem. Soc., 2010, 132, 14385.

6782 Chem. Commun., 2011, 47, 6763–6783 This journal is c The Royal Society of Chemistry 2011
View Article Online

180 R. I. Bickley, T. Gonzalezcarreno, J. S. Lees, L. Palmisano and 195 I. Nakamura, N. Negishi, S. Kutsuna, T. Ihara, S. Sugihara and
R. J. D. Tilley, J. Solid State Chem., 1991, 92, 178. E. Takeuchi, J. Mol. Catal. A: Chem., 2000, 161, 205.
181 D. C. Hurum, A. G. Agrios, K. A. Gray, T. Rajh and 196 J. Woning and R. A. Vansanten, Chem. Phys. Lett., 1983, 101, 541.
M. C. Thurnauer, J. Phys. Chem. B, 2003, 107, 4545. 197 M. Lazzeri and A. Selloni, Phys. Rev. Lett., 2001, 87, 266105.
Published on 29 March 2011. Downloaded by Universidade Federal do Espirito Santo (UFES) on 4/22/2020 12:51:15 AM.

182 D. C. Hurum, K. A. Gray, T. Rajh and M. C. Thurnauer, J. Phys. 198 G. S. Herman, M. R. Sievers and Y. Gao, Phys. Rev. Lett., 2000,
Chem. B, 2005, 109, 977. 84, 3354.
183 T. Kawahara, Y. Konishi, H. Tada, N. Tohge, J. Nishii and 199 R. Hengerer, B. Bolliger, M. Erbudak and M. Gratzel, Surf. Sci.,
S. Ito, Angew. Chem., Int. Ed., 2002, 41, 2811. 2000, 460, 162.
184 J. Zhang, Q. Xu, Z. Feng, M. Li and C. Li, Angew. Chem., Int. 200 N. Ruzycki, G. S. Herman, L. A. Boatner and U. Diebold, Surf.
Ed., 2008, 47, 1766. Sci., 2003, 529, L239.
185 X. W. Wang, G. Liu, Z. G. Chen, F. Li, L. Z. Wang, G. Q. Lu 201 L. Liu, J. S. Qian, B. Li, Y. M. Cui, X. F. Zhou, X. F. Guo and
and H. M. Cheng, Chem. Commun., 2009, 3452. W. P. Ding, Chem. Commun., 2010, 46, 2402.
186 X. W. Wang, G. Liu, G. Q. Lu and H. M. Cheng, Int. J. Hydrogen 202 O. Bikondoa, C. L. Pang, R. Ithnin, C. A. Muryn, H. Onishi and
Energy, 2010, 35, 8199. G. Thornton, Nat. Mater., 2006, 5, 189.
187 H. M. Chen, C. K. Chen, Y. C. Chang, C. W. Tsai, R. S. Liu, 203 K. Onda, B. Li, J. Zhao, K. D. Jordan, J. L. Yang and H. Petek,
S. F. Hu, W. S. Chang and K. H. Chen, Angew. Chem., Int. Ed., Science, 2005, 308, 1154.
2010, 49, 5966. 204 A. Ignatchenko, D. G. Nealon, R. Dushane and K. Humphries,
188 M. T. Niu, F. Huang, L. F. Cui, P. Huang, Y. L. Yu and J. Mol. Catal. A: Chem., 2006, 256, 57.
Y. S. Wang, ACS Nano, 2010, 4, 681. 205 A. Tilocca and A. Selloni, J. Phys. Chem. B, 2004, 108, 4743.
189 R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki and Y. Taga, 206 E. M. J. Johansson, S. Plogmaker, L. E. Walle, R. Scholin,
Science, 2001, 293, 269. A. Borg, A. Sandell and H. Rensmo, J. Phys. Chem. C, 2010,
190 S. U. M. Khan, M. Al-Shahry and W. B. Ingler, Science, 2002, 114, 15015.
297, 2243. 207 C. G. Read, E. M. P. Steinmiller and K. S. Choi, J. Am. Chem.
191 X. T. Hong, Z. P. Wang, W. M. Cai, F. Lu, J. Zhang, Y. Z. Yang, Soc., 2009, 131, 12040.
N. Ma and Y. J. Liu, Chem. Mater., 2005, 17, 1548. 208 T. Ohno, K. Sarukawa and M. Matsumura, New J. Chem., 2002,
192 G. Liu, Z. G. Chen, C. L. Dong, Y. N. Zhao, F. Li, G. Q. Lu and 26, 1167.
H. M. Cheng, J. Phys. Chem. B, 2006, 110, 20823. 209 S. V. Yanina and K. M. Rosso, Science, 2008, 320, 218.
193 T. Umebayashi, T. Yamaki, H. Itoh and K. Asai, Appl. Phys. 210 T. Tachikawa, N. Wang, S. Yamashita, S. C. Cui and T. Majima,
Lett., 2002, 81, 454. Angew. Chem., Int. Ed., 2010, 49, 8593.
194 W. Ho, J. C. Yu and S. Lee, Chem. Commun., 2006, 1115. 211 T. Tachikawa and T. Majima, Chem. Soc. Rev., 2010, 39, 4802.

This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 6763–6783 6783

You might also like