Sie sind auf Seite 1von 7

MSE 523 Fall 2015

Project
Mini Review Article
Attach this cover sheet to the front of mini review article when you submit it.

Name: Diane Haiber


ASU ID: 1203754209

Title of Article: Effects of Crystal Structure in Metal-Oxide Semiconductors on Photocatalytic Activity for Water
Splitting

Abstract: The interplay between the crystal structure and electronic structure in metal oxide semiconductors will
be emphasized through three themes as they relate to photocatalytic water splitting: (i) charge transport pathways,
(ii) characteristics of layered perovskites, and (iii) active surfaces on nanocrystals. Concepts of metal-oxide
configurations within the unit cell, total cation valency, metal-oxygen bonding angles, and perovskite layer
thickness are discussed. Approaches to studying the activities of crystallographic surfaces in titania, niobates, and
gallium-oxides involve first-principle studies. Understanding structure-reactivity relationships in photocatalytic
systems will be accelerated by use of high-resolution in-situ and ex-situ electron microscopy.

Grading rubric:

Criterion
Topic and scope
Technical content
Use of references
Use of figures
Analysis and discussion
Organization and clarity

Grade: ____________ / 7

Excellent Good Fair Poor

EFFECTS OF CRYSTAL STRUCTURE IN METAL-OXIDE SEMICONDUCTORS ON PHOTOCATALYTIC


ACTIVITY FOR WATER SPLITTING
INTRODUCTION: As the carbon footprint associated with fossil fuel energy production continues to grow,
future generations will need alternative, environmentally-benign systems to meet their energy needs. An
energy economy based on hydrogen has the potential to store solar energy and limit greenhouse gas
emission. Most hydrogen gas is obtained from carbon-emitting processes such as reforming of fossil fuel.16
Enabling the reaction H2OO2+H2 with sunlight on semiconducting materials is the basis for
photocatalytic water splitting wherein solar energy is converted into chemical energy. This reaction depends
upon the semiconductors ability to absorb photons to generate electron/hole pairs of sufficient energy (at
least 2 eV is needed to kinetically drive water splitting16), and diffusion of these charges to surface active
sites where redox reactions occur.
Crystal structure has a significant effect on the ability of
a material to perform as a photocatalyst because it is directly
related to the electronic structure (e.g. electrical conductivity).
Defects associated with grain boundaries can act as charge
recombination sites, which will lower the activity. Figure 1
illustrates the effect that crystal structure of TiO2 has on its
electronic structure.14 In the amorphous phase where there is no
long range order, the crystallinity is low and there is
correspondingly low activity due to low carrier mobility and
increased charge recombination rates. Upon calcination, or
heating in vacuum, the amorphous TiO2 crystallizes to the
Figure 1: Crystallinity and surface area of
anatase phase and further to the rutile phase wherein the
different phases of TiO2.14
photocatalytic activity becomes highest for rutile (highest
crystallinity). Both anatase and rutile are in the tetragonal system and have the same composition. However,
rutile has a smaller band gap energy of 3 eV than that of anatase at 3.2 eV this suggests that the crystal
structure directly influences the electronic structure of a semiconductor. Note that surface area also
has an opposing effect on photocatalytic activity through the decrease of active sites but this factor will not
be discussed. Surface area generally does not affect photocatalytic activity as significantly as crystallinity
does due to the charge recombination effect being the limiting step.
To experimentally study the effects of crystal structure on photocatalytic activity, both water splitting
reactions and certain materials characterization techniques must be employed. X-ray diffraction (XRD) is
one of the most common approaches for determination of the crystal structure of a material through
characterization of lattice spacing in a bulk crystal or powder. It can also be used to detect changes in
crystallinity via differences in the half-width of specific peaks, which is particularly useful for monitoring
variations in crystallinity between materials of the same crystal structure. Ultraviolet-visible (UV/Vis)
absorbance or diffuse reflection measurements are used to determine the band gap energy of the bulk
material. More advanced characterization techniques include scanning electron microscopy (SEM) and
transmission electron microscopy (TEM) are used to measure particle size distribution/morphology and
local crystal and electronic structures.
The interplay between the crystal structure and electronic structure in metal oxide semiconductors will
be emphasized through three themes as they relate to photocatalytic water splitting:
(i) charge transport pathways in semiconductors
(ii) characteristics of layered perovskites, and
(iii) active surfaces on nanocrystals.
The five main articles that highlight themes (i)-(iii) include discussions on metal-oxide octahedron
configurations within the unit cell, total cation valency in layered perovskites, and active surfaces of
nanocrystals.1-5 This review will not focus on the effects of co-catalyst loading on photocatalyst materials.
Most articles report the optimal conditions achieved through co-catalyst loading but these discussions will
be omitted as they do not directly relate to the crystal structure of the photocatalyst.
Haiber, 1

active

CHARGE TRANSPORT PATHWAYS IN SEMICONDUCTORS: In a systematic


analysis of the crystal structures and photocatalytic activities of metal
oxides of different constituent elements, R3MO7 and R2Ti2O7 (R = Y, Gd,
La; M = Nb, Ta), it was found that water splitting was associated with the
orthorhombic weberite, cubic pyrochlore (only for R2Ti2O7 group), and
monoclinic perovskite structure.1 Orthorhombic weberite, which was
demonstrated by calcined La2TaO7 and La3NbO7, is characterized by corner
sharing metal-oxide octahedron chains that zig-zag across the {001} planes.
inactive
The La3+ ions sit between the octahedron. Cubic pyrochlore also forms
(b)
these octahedron chains, but the titanates (R2Ti2O7) were the only types that
showed activity with this crystal structure. Upon further inspection, it can
be seen that in the R3MO7 group of cubic pyrochlores, two types of metaloxide octahedron can form: RO6 and MO6, which results in a mixed network
of metal-oxide octahedron. In Figure 2 the inactive and active types of cubic
pyrochlore structures are illustrated. Thus, the titanates form homogenous
corner sharing octahedron whereas the second group forms an alternating
Figure 2: Schematic diagram of pattern (e.g. TaO6 and GdO6 in Gd3TaO7). As a result, it was concluded that
the unit cells of cubic pyrochlore conductive pathways are not formed in the latter leading to photocatalytic
structure for (a) Y3Ti2O7 and (b)
inactivity.
Gd3TaO7.1
It should also be noted that the arrangement of atoms within the unit cell
for the R3MO7 materials with the cubic pyrochlore structure (represented in Fig. 2b) was constructed by
assuming alternating octahedral units of 1:1 ratio.1 This is a reasonable assumption given the XRD data
confirming a cubic pyrochlore structure and the assumption that is material will be similar to the general
A2B2O7 cubic pyrochlore (Fig.2a). A more complete analysis may include high angle annular dark field
scanning transmission electron microscopy of a material with the proposed mixed metal cation octahedron
structure, which would allow for direct observation of the atomic columns through Z-contrast imaging.
Kato and Kudo have found that stoichiometric alkali
tantalates, ATaO3 (A = Li, Na, and K), possess different
photocatalytic activities due to the effect that crystal structure
has on electronic properties.2 As the alkali metal ion radius
decreases, the bond angle deviates from 180. Figure 2
illustrates this effect as K, Na, and Li have O-Ta-O bonding
angles of 180, 163, and 143, respectively. The band gap
energies for K-, Na-, and Li- based tantalates are 3.6, 4.0, and
4.7 eV. Thus, decreased optical absorption is associated with
distorted bond angles. As a result, photocatalytic activity for
KTaO3 is highest. To contrast previous points regarding
photocatalytic activity in homogeneous corner-sharing
octahedron, this report finds that the orientation of adjacent
octahedron is also an important factor in determining activity. Figure 3: Schematic diagram ATaO3 (A=Li, Na,
2,14
To explain the effect of TaO6 octahedron bonding angles K), important characteristics, and activities.
on the electronic structure, which in turn will affect the photocatalytic activity, Wiegel et al. previously
studied the luminescent properties of the same three materials as above.7 A comparison between the
unknown luminescence behavior in ATaO3 and well-characterized ANbO3s (A = Li, Na, and K) confirmed
that the O-Ta-O bonding angle favored charge delocalization as it approached 180 where pi-bonds are
easily formed. Therefore, charge transport more easily occurs in KTaO3 due to the increased delocalization
facilitated by its structure.
In M2Ti6O13 (M = Na, K, and Rb) studied by Ogura et al., rectangular tunnel structures in TiO6 cornersharing octahedrons resulting from unit cell dimensional variations were found to have a significant impact
on photocatalytic activity.6 As the size of M increases, the unit cell dimensions a and b decrease while c
extends and this distortion gives rise to dipole moments within the metal-oxide octahedrons. Water splitting
(a)

Haiber, 2

activity was correlated to the strength of the dipole moment which was greatest for the smallest ion, Na. It
was concluded that the dipole moment supported photoexcited charge separation, which lead to increased
activity. Tokunaga et al. studied the differences in BiVO4 scheelite structures with tetragonal and
monoclinic symmetry.9 While both materials were found to have the same band gap energy through UV/Vis
diffuse reflection spectra, oxygen evolution was much higher in the monoclinic structure (about 600 to ~0
mol hr-1 under visible light). Between the tetragonal and monoclinic structure, the variations in the a, b,
and c cell parameters and angle are indicative of distorted bismuth oxide polyhedron, which will effect
charge separation efficiency of electrons and holes.
LAYERED PEROVSKITE PHOTOCATALYSTS: Layered
perovskites have become popular in terms of photocatalytic
water splitting due to their ability to capture water in the
interlayer space which reduces the charge migration
distance necessary for reactions to occur.15 In layered
perovskites materials, the interlayer space can be parallel to
different miller planes and hence are referred to in terms of
these indices. Figure 4 illustrates three types of layered
perovskite structures in terms of their relationship to the
boundaries of the metal-oxide octahedron.3 In a study by
Miseki, Kato, and Kudo, the photocatalytic activity of
(111)-type layered perovskites were studied.3 The materials Figure 4: Different types of layered3perovskites,
viewed in the a-c plane.
were metal oxides with the stoichiometric formulas
A5Nb4O15 (A = Sr and Ba), Ba3LaNb3O12, ALa4Ti4O15 (A = Ca, Sr, and Ba), and La4Ti3O12. Using XRD
and UV/Vis diffuse reflectance spectra, it was found that the band gap energy widened as the bonding angle
between metal and oxygen ions diverged from 180, similar to findings in other metal oxide materials.2,7
For niobates Sr5Nb4O15 and Ba5Nb4O15, where Sr and Ba have different ionic radii, the bonding angles are
168 and 180, respectively, and correspond to band gap energies of 4.04 and 3.91 eV. Conversely, the band
gap for Ba3LaNb3O12 (three-octahedron layer thickness) was 4.1 eV whereas narrower band gaps for the
four-layered A5Nb4O15 (A = Sr, Ba) and ALa4Ti4O15 (A = Ca, Ba) materials ranged from 3.79-4.04 eV. In
this study, layer thickness did effect the photocatalytic activity by increasing the excited-state localization
and widening the band gap, which leads to decreased photogenerated electrons and holes. Lastly, the
distance between adjacent metal-oxide octahedron layers, which is dictated by the ion radius that sits
between the layers, was found to effect the band gap energy and corresponding photocatalytic activity. This
was demonstrated for Ba3LaNb3O12 and CaLa4Ti4O15, where a wider bang gap was associated with Bacontaining material since Ba2+ is larger than Ca2+.
The (110) layered perovskites were experimentally found to be superior to the (100) type in a study by
Kim and Hwang et al.18 Samples under study consisted of AmBmO3m+1 (m=4, 5; A = Ca, Sr, La; B = Nb, Ti)
layered perovskites that are highly donor doped; their overall water splitting activity was compared to
commercial grade TiO2 and two (100) type layered perovskites. The highest quantum yield of 23% (i.e.
percentage of absorbed photons used to create reaction products) was found for Sr2Nb2O7. Due to the cation
valency of Nb5+ in this material relative to the valency of Ti4+ in an ideal perovskite such as SrTiO3, resulting
excess electrons are is the source of the so-called donor doping.
Kim et al. studied (110) layered perovskites layer thickness and total cation valency on their
photocatalytic properties in titanates and niobates.8 Interestingly, there was no correlation between the layer
thickness and catalytic activity. The water reduction reaction was monitored over three LaxTiyOz layered
perovskites with slabs composed of 1, 3, and 4 TiO6 units; the activities were 442, 386, and 474 mol hr1 -1
g . Total cation valency is determined by the constituent elements and the layer thickness. This was the
primary factor for photocatalytic activity due to the enhanced electron-hole separation achieved over the
thinner depletion layer. The Sr-O-Nb, Sr-O-Ta and La-O-Ti layered structures have +7 valency and
hydrogen evolution activities from 402-796 mol hr-1g-1 while Sr-O-Ti of +6 valency has a very low activity
of 11 mol hr-1g-1.
Haiber, 3

To enhance the visible light absorption through narrowing of the band gap, Sun et al. doped strontium
titanates (Sr2TiO4) of (001) layer type with the transition metal Cr.10 Doping refers to intentionally adding
small concentrations of foreign elemental species such that the bulk crystal structure is unchanged. This
results in electron deficient or electron excess regions within the bulk structure that result in improved
electrical conductivity. XRD analysis identified Sr2TiO4 has a tetragonal crystal structure with I4/mmm
symmetry; upon doping with different amounts of Cr the a and c dimensions of the unit cells were
decreased. Moreover, UV/Vis analysis showed decreased absorption edge for 0.60 and 0.33 Cr mol %
doped strontium titanates, corresponding to increased visible light absorption. The highest photocatalytic
activity was demonstrated by 0.60 Cr mol% doped Sr2TiO4, with a hydrogen evolution rate of 977 mol g1 -1
hr .
Kato, Kudo, and Nakagawa studied stronium niobates and tantalates (Sr2Ta2O7 and Sr2Nb2O7) to
explore if differences in crystal structure or electronic structure are the main factor in determining the
activities between these types of layered perovskites.13 In this case, Sr2Ta2O7 was the more active
photocatalyst with H2 and O2 evolution rates of 52 and 18 mol g-1hr-1, respectively, but also had the larger
band gap energy. Sr2Nb2O7 evolved no oxygen and evolved hydrogen at the rate of 5.9 mol g-1hr-1 but has
a smaller band gap of 3.9 eV. Generally, a narrowed band gap means increased photon absorption and
photogenerated electron-hole pairs, which increases the overall water splitting efficiency. The importance
of energy band alignment is emphasized here. The conduction band minimum for Sr2Ta2O7 has more
negative electrical potential than Sr2Nb2O7s, meaning there is an increased overpotential to drive the water
reduction process in the tantalate material. Because these crystal structures of these two materials are
similar, it was concluded that the primary factor in determining the differences in their activities is the
conduction band energy alignment.
ACTIVE CRYSTALLOGRAPHIC SURFACE FACETS: The
success of photocatalytic reactions, namely water splitting, is
determined by the surface it is being facilitated on. Hence,
much research has been focused on identifying highly active
surfaces of different materials so that more efficient
photocatalysts can be synthesized. Zhou et al. experimentally
determined that the {111} family of planes are the most active
for photocatalytic water splitting in the NiGa2O4 system by
comparing the hydrogen evolution rates of octahedral
nanocrystals and nanorods.4 Synthesizing octahedron
Figure 5: Photocatalytic activity of NiGa2O4
nanocrystals with exposed {111} faces and nanorods grown
octahedrons and nanorods, representing the
in the [100] direction allowed assessment of the activities of
{111} and {001} surfaces, respectively.4
different crystallographic surfaces in NiGa2O4. Figure 5
shows the water splitting activities of both sample geometries. Hydrogen evolution was highest for all
octahedral nanocrystals and lowest for the nanorods, even though the nanorods possessed a 3x greater
surface area. This confirms that the surface structure is most important. Upon density functional theory
calculation, it was found that a {111} surface, composed of mixed 4-fold and 6-fold coordinated Ga and Ni
atoms, had the lowest surface energy and is therefore is most likely to form. This structure of mixed
statistical coordination, relative to other candidates, provides more Ga or Ni atoms at the surface, which
means there an increased number of active sites. Transient photocurrent decay, a process in which the
photocurrent is measured during light-on intervals, was used to asses charge recombination rates on the two
surfaces and found the {111} surface to have a lower recombination rate owing to its high activity.
To understand the structure-reactivity relationship in potassium niobates (KNbO3), nanocrystals with
tetragonal and orthorhombic structures were synthesized and compared to data on those with cubic
structure.11 XRD, Raman spectroscopy, and X-ray photoelectron spectroscopy were able to reveal the minor
differences in structure between the three phases. The band gap energies for cubic, orthorhombic, tetragonal
are 3.24, 3.15, and 3.18 eV, respectively. In terms of symmetry, the cubic, orthorhombic, and tetragonal
phases have Oh, C2v, and C4v point group symmetries for their unit cells. It is generally accepted that the
high symmetry is related to high photocatalytic activity which is consistent with cubic KNbO3 having the
Haiber, 4

highest activity. Upon more detailed computational analysis, modelling of the lowest unoccupied wave
functions (LUWF) for the orthorhombic and tetragonal phases revealed that the orthorhombic LUWF is
preferentially located near the surface and subsurface layers. This means there is a higher probability of
photo generated electrons residing near the surface active sites, which explains its higher activity over the
tetragonal phase. Moreover, the effect of exposed crystallographic surfaces on photocatalytic activity was
emphasized.
Titania (TiO2) was the first material
discovered as a photocatalyst for solar water
splitting and there has been a lot of debate over
which is the most active surface. Pan et al.
described their finding that the order of activity
in low-index facets for TiO2 anatase nanocrystals
is {010}>{101}>{001}, which was based on
photoreaction experiments on well-defined
samples.5 Figure 6 illustrates the relative surface
areas of such nanocrystals composed of these
low-index planes both schematically and
experimentally. Interestingly, the {001} facet has
Figure 6: Diagrams (a) and corresponding SEM images the highest surface density of undercoordinated
(b-d) of preferentially faceted nanocrystals of TiO2.5
Ti atoms but shows the lowest activity. Upon
surface electronic structure analysis via UV/Vis spectroscopy and XPS, it was found that the conduction
band (CB) minimum for {001} materials is downshifted to more positive electrical potential, leading to a
decreased overpotential for hydrogen reduction. For all three phases, the valence band alignment remained
unchanged. Facet-dependent effects were eliminated by terminating the surfaces with fluorine, which
suggests the coordination of surface Ti atoms is a significant factor in determining the activity. Zhao et al.
developed a simple technique for synthesizing stable highly active anatase TiO2 nanocrystals enclosed
by the {001} facets in the same year as the previous study.12 Although titania is the most widely studied
photocatalytic system, there is still much debate over the fundamentals of its structure and reactivity
relationships.
CONCLUSION: Metal-oxide semiconductors represent a large percentage of photocatalytic materials for
water splitting being studied today. Their structures are based on corner-sharing metal-oxide octahedrons
in which the constituent atoms and stoichiometry dictate finer features such as crystal structure, metal-oxide
octahedron orientation, compositional distribution, and layered-perovskite character. Variations in the
atomic-scale structure are directly related to the electronic structure and resulting ability of the material to
generate electron-hole pairs and allow them to diffuse to surface active sites. The importance of charge
transport pathways in titanate, niobate, and tantalates was defined by delocalized excited energy states,
which were enabled by homogeneous corner-sharing octahedrons, non-distorted bonding, and internal
space charge regions that support charge separation.1,2,6,7,9 Another important class of metal-oxides are
layered perovskite structures since their interlayer space can be incorporated with water to boost
photocatalytic activity. Similar factors dictate their ability to facilitate fast charge separation and transfer
but layer thickness and crystal structure contributions seem to change depending on the orientation of the
layers.3,8,13 Much work has also been carried out to understand the activities of certain crystallographic
surfaces, as catalytic water splitting is a surface-mediated reaction. Experimental observations of carefully
synthesized nanocrystals together with first principle models confirm which surfaces are most efficient for
photocatalytically splitting water.4,11 While many characterization techniques such as XRD and UV/Vis
probe the bulk structure, nano-scale observations of local electronic and atomic structures of photocatalysts
during reaction conditions are also highly informative for understanding structure-reactivity
relationships.18,20 Future researchers will continue to discover new photocatalyst materials. On the other
hand, in-situ environmental transmission electron microscopy studies will deepen knowledge and possible
deactivation mechanisms of well-defined photocatalytic systems by gaining morphological, chemical,
electronic, and structural information during reaction conditions.
Haiber, 5

REFERENCES
[1] Abe, R.; Higashi, M.; Sayama, K.; Abe, Y.; Sugihara, H. Photocatalytic Activity of R 3MO7 and
R2Ti2O7 (R= Y, Gd, La; M = Nb, Ta) for Water Splitting into H2 and O2. J. Phys. Chem. B 2006,
110, 2219-2226.
[2] Kato, H.; Kudo, A. Water Splitting into H2 and O2 on Alkali Tantalate Photocatalysts ATaO3 (A =
Li, Na, K). J. Phys. Chem. B 2001, 105, 4285-4292.
[3] Miseki, Y.; Kato, H.; Kudo, A. Water Splitting into H2 and O2 over Niobate and Titanate
Photocatalysts with (111) Plane-Type Layered Perovskite Structure. Energy Environ. Sci. 2009, 2,
306-314.
[4] Zhou, S.; Lv, X.; Zhang, C.; Huang, X.; Kang, L.; Lin, Z.; Chen, Y.; Fu, W. Synthesis of NiGa2O4
Octahedron Nanocrystal with Exposed {111} Facets and Enhanced Efficiency of Photocatalytic
Water Splitting. ChemPlusChem 2015, 80, 223-230.
[5] Pan, J; Liu, G.; Lu, G.; Cheng, H. On the True Photoreactivity Order of {011}, {010}, and {101}
Facets of Anatase TiO2 Crystals. Angew. Chem. Int. Ed. 2011, 50, 2133-2137.
[6] Ogura, S.; Kohno, M.; Sato, K.; Inoue, Y. Photocatalytic Activity for Water Decomposition of
RuO2-combined M2Ti6O13 (M = Na, Rb, Cs). Appl. Surf. Sci. 1997, 121, 521-524. (Impact factor =
2.711)
[7] Wiegel, M.; Emond, M.; Stobbe, E.; Blasse, G. Luminescence of Alkali Tantalates and Niobates.
J. Phys. Chem. Solids 1994 55 (8), 773-778.
[8] Kim, J.; Hwang, D.; Kim, H.; Bae, S.; Lee, J.; Li, W., Oh, S. Highly Efficient Overall Water
Splitting Through Optimization of Preparation and Operation Conditions of Layered Perovskite
Photocatalysts. Top. Catal. 2005, 35 (3-4), 295-303.
[9] Tokunaga, S.; Kato, H.; Kudo, A. Selective Preparation of Monoclinic and Tetragonal BiVO4 with
Scheelite Structure and their Photocatalytic Properties. Chem. Mater. 2001, 13, 4642-4628.
[10] Sun, X.; Xie, Y.; Wu, F.; Chen, H.; Lv, M.; Ni, S.; Liu, G., Xu, X. Photocatalytic Hydrogen
Production over Chromium Doped Layered Perovskite Sr2TiO4. Inorg. Chem. 2015, 54, 7445-7453.
[11] Zhang, T.; Zhao, K.; Yu, J.; Qi, Y.; Li, H.; X. Hou; Liu, G. Photocatalytic Water Splitting for
Hydrogen Generation on Cubic, Orthorhombic, and Tetragonal KNbO3 Microcubes. Nanoscale
2013, 5, 8375-8383.
[12] Zhao, X.; Jin, W.; Cai, J.; Ye, J.; Li, Z.; Ma, Y.; Xie, J.; Qi, L. Shape- and Size-Controlled Synthesis
of Uniform Anatase TiO2 Nanocuboids Enclosed by Active {100} and {001} Facets. Adv. Funct.
Mater. 2011, 21, 33445-3563.
[13] Kudo, A.; Kato, H.; Nakagawa, S. Water Splitting into H2 and O2 on New Sr2M2O7 (M = Nb and
Ta) Photocatalysts with Layered Perovskie Structures: Factors Affecting the Photocatalytic
Activity. J. Phys. Chem. B 2000, 104, 571-575.
[14] Kudo, A.; Miseki, Heterogeneous Photocatalyst Materials for Water Splitting. Chem. Soc. Rev.
2009, 38, 253-278.
[15] Osterloh., F. Inorganic Materials as Catalysts for Photochemical Splitting of Water. Chem. Mater.
2008, 20, 35-54.
[16] Yergo, R.; Galvn, M.; del Valle, F.; de la Mona, J.; Fierro, J. Water Splitting on Semiconductor
Catalysts under Visible-Light Irradiation. ChemSusChem 2009, 2, 471-485.
[17] Ohno, T.; Sarukawa, K.; Matsumura, M. Crystal Faces of Rutile and Anatase TiO2 Particles and
their Roles in Photocatalytic Reactions. New J. Chem. 2002, 26, 1167-1170.
[18] Kim, H.G.; Hwang, D.W.; Kim, J.; Kim, Y.G.; Lee, J.S. Highly donor-doped (110) layered
perovskite materials as novel photocatalysts for overall water splitting. Chem. Commun. 1999,
1077-1078.
[19] Zhang, L.; Miller, B.K.; Crozier, P.A. Atomic Level In Situ Observation of Surface Amorphization
in Anatase Nanocrystals During Light Irradation in Water Vapor. Nano Lett. 2013, 13, 679-684.
[20] Liu, Q.; Zhang, L.; Crozier, P.A. Strucutre-reactivity relationships of Ni-NiO core-shell cocatalysts on Ta2O5 for solar hydrogen production. Applied Catalysis B: Environmental 2015, 172173, 58-64.
Haiber, 6

Das könnte Ihnen auch gefallen