Applied Catalysis A: General: Gerardo Colón

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Applied Catalysis A: General 518 (2016) 48–59

Contents lists available at ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Towards the hydrogen production by photocatalysis


Gerardo Colón
Instituto de Ciencia de Materiales de Sevilla, Centro Mixto CSIC-Universidad de Sevilla, C/Américo Vespucio, 49 41092 Sevilla, Spain

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

Article history: Nowadays, problems derived from climate change urgently demand us to focus our attention on new
Received 23 July 2015 alternatives to fossil fuels. Within this framework, the photocatalytic production of hydrogen as a clean
Received in revised form fuel from oxygenates arises as a necessary option that must be considered. Thus, the development of
23 November 2015
highly efficient photocatalyst is crucial in order to achieve a viable technology under the industrial point
Accepted 24 November 2015
of view. For this sake, it is necessary to understand the principles of photoreforming reaction. In this brief
Available online 30 November 2015
review we will revisit the different photocatalytic materials proposed in the literature highlighting on
the role of different co-catalysts.
Keywords:
Photocatalysis © 2015 Elsevier B.V. All rights reserved.
Hydrogen production
Co-catalysts

1. Introduction ment, and offers unique opportunities, benefits and challenges. At


present, H2 is mainly produced from CO and CH4 from fossil fuels by
In the early 1874, the visionary writer Jules Verne predicted that a steam reforming reaction which is accomplished by coupling with
“water will be the coal of the future” in his science fiction tale “The water gas shift and hydrogen purification reactions [3]. However,
Mistery Island”. A century later, Fujishima and Honda could state in the forthcoming effective reduction of fossil fuel reserves as well as
the lab the prediction made by Verne [1]. In this widely referenced the serious environmental problems associated with CO2 produc-
experiment they showed that band gap excitation of anatase TiO2 tion has inspired the development of viable alternatives. Among
in a photoelectrochemical cell with a Pt counter electrode and an these, the solar photocatalytic application for hydrogen produc-
applied bias resulted in water splitting into hydrogen and oxygen. tion by photoreforming in which sunlight and water are used as
This scientific result has becoming the inspiring step in the devel- the hydrogen source is a highly appreciated alternative [4]. In spite
opment of a large studied field of the photocatalysis focusing on of these attractive particularities, the actual hydrogen production
the water splitting reaction. Although the extensive research over rates are still far from a practical application status. Thus, several
the years in order to refine the process, nobody has come up with reported works on high scale reactors indicated that the obtained
a system that is both efficient and inexpensive enough to produce H2 productions are quite modest [5]. It is clear that the optimiza-
sufficient hydrogen to be used as a clean-burning fuel on the roads, tion of the photocatalytic system is crucial in order to make this
in industry, and at home. attractive technology feasible under the industrial point of view.
The relevancy of this discovery is getting more evident taking In this sense, Rodríguez et al. recently showed that hydro-
into consideration the socioeconomical and environmental situa- gen generated from solar sources could be competitive with that
tion in the last decades. It is widely believed that hydrogen will obtained from non-renewables, and point towards design strate-
play an important role in this global system since it is considered gies that can significantly aid in the reduction of solar-fuels
the ultimate clean energy carrier. Hydrogen can be produced from production costs [6]. Similarly, Pinaud et al. clearly demonstrates
a variety of feedstocks. These include fossil resources, such as nat- that if technical progress is made to meet material performance
ural gas and coal, as well as renewable resources, such as biomass targets and with appropriate plant-scale engineering, direct solar
and water with input from renewable energy sources (e.g. sunlight, hydrogen produced by water splitting reaction can be produced at
wind, wave or hydro-power). A variety of process technologies can a cost which meets the US Department of Energy target thresh-
be used, including chemical, biological, electrolytic, photolytic and old of $2.00–$4.00 per kg H2 [7]. These authors performed an
thermo-chemical [2]. Each one is in a different stage of develop- exhaustive analysis which revealed that improvement in the
solar-to-hydrogen efficiency (STH, which correlates the hydrogen
production rate and energy of incidence solar light) of the panel-
based systems could substantially drive down their costs. For a
E-mail address: gcolon@icmse.csic.es

http://dx.doi.org/10.1016/j.apcata.2015.11.042
0926-860X/© 2015 Elsevier B.V. All rights reserved.
G. Colón / Applied Catalysis A: General 518 (2016) 48–59 49

Fig. 1. Bandgaps and band edge positions of selected semiconductors in relation to the redox potentials for water splitting reaction.

given STH of 10% and a photocatalyst lifetime of ten years, the on actually understanding the process itself. Recent advances in
price of hydrogen was estimated to be $1.6 per kg, which could the tailoring of new photocatalysts for solar water splitting pass
meet the mentioned target hydrogen price. In a recent perspective through the comprehension of the band electronic structure which
analysis, Domen et al. argued that due to the limitation of UV- subsequently would lead to improvements using band engineering.
active photocatalyst to increase the STH, it is necessary to develop
and activate narrow-band-gap semiconducting photocatalysts for 2. General considerations on H2 production reaction
practical operation despite the present low activity of photocata-
lysts with absorption edge wavelengths longer than 600 nm for the From a thermodynamic point of view the water splitting
overall water-splitting reaction [8]. These authors finally conclude reaction is energetically unfavoured process. Thus, water trans-
that photocatalytic systems must be designed bearing scalability in formation into H2 and O2 is an uphill reaction which needs the
mind. standard Gibbs free energy change of 237 kJ/mol.
Over the last three decades the overall photocatalytic water
splitting process has been extensively studied pursuing to enhance 2 H+ + 2e− → H2 Hydrogenevolutionhalfreaction
the photoefficiency of the process [9]. However, the earliest
research focused more on developing good photocatalysts than 2OH− + 2 h+ → ½O2 + H2 OOxygenevolutionhalfreaction

H2 O → H2 + ½O2 Overallwatersplitting

The redox potentials for both reduction and oxidation processes


determine the possible candidate photocatalyst by the positions
of the valence and conduction bands and therefore strongly limit
the range of possible photocatalysts. Both the reduction and oxi-
dation potentials of water should lie within the band gap of the
photocatalyst.
Regarding to the formal electronic structure of a potential pho-
tocatalyst for H2 production, there are two important requirements
to be accomplished: (i) The band gap should be 1.23 eV < Eg <
3.26 eV. (ii) The band positions should be located as follows: the
bottom of the conduction band should be more negative than the
redox potential of H+ /H2 (0 V vs NHE) meanwhile the top of the
valence band should be more positive than the redox potential of
O2 /H2 O (1.23 V).
Thus, from a thermodynamic point of view the water splitting
reaction should be easily achieved using a photoinduced catalytic
process involving any materials which satisfy the above conditions.
However, the overall photocatalytic water splitting reaction is an
endothermic reaction. This means that, in order to overcome such
an energy barrier, photons of higher energy are needed.
Within the great number of photoactive semiconductors pro-
Fig. 2. Morphology of anatase nanocrystals. posed for water splitting ans photoreforming reactions, till now
50 G. Colón / Applied Catalysis A: General 518 (2016) 48–59

Fig. 3. Possible homojunction formation by TiO2 polymorphs.

TiO2 is the most efficient one. However, the reported photonic be increased at the same time as efficiency. On this basis, pho-
efficiencies from solar to hydrogen by TiO2 photocatalytic water- toreforming can be considered a more feasible process than water
splitting are still low. This is mainly due to the following reasons: splitting under the thermodynamic point of view [19].
(1) Recombination of photo-generated electron/hole pairs: con- If we consider the H2 evolution half reaction by means of a pho-
duction band electrons can recombine with valence band holes toreforming process, a generally accepted mechanism involves the
very quickly and release energy in the form of unproductive heat oxidation of water molecules by photoinduced holes in the semi-
or photons; (2) Competitive fast backward reaction: As mentioned conductor. In this step, OH• radicals can be produced which can
before the water cleavage into hydrogen and oxygen is an energy attack the alcohol (acting as sacrificial agent in the oxidation half
unfavoured process, thus backward reaction could easily proceeds reaction) and abstract an alpha hydrogen to form a • RCH2 -OH rad-
and strongly competes with water splitting; (3) Failure to absorb ical [20].
visible light photons: The band gap of TiO2 is about 3.2 eV and Photoreforming was initially studied by using methanol, the
only UV light can be utilized for hydrogen production. Since the simplest molecule, however in recent papers biomass derived
UV light only accounts for about 4% of the solar radiation energy, compounds were considered, providing this process an additional
the efficiency of process becomes drastically low. interest. Thus, H2 were obtained from ethanol, [21,22]. glycerol
In order to resolve the above mentioned handicaps different [23,24], glucose [25,26], sucrose [27], and other biomass derived
approaches have been proposed in the literature. compounds [28].
Thus, the recombination of the photogenerated charge carriers In a recent paper Bahruji et al. studied in detail the H2 production
has been afforded since the starting research works are mainly reaction by photoreforming of different alcohols and proposed a
focused on TiO2 based catalysts [10]. Indeed, for pure TiO2 , non- chemical mechanism of H2 production reaction [29]. These authors
radiative recombination appears to be the main de-excitation proposed interesting rules that could govern the alcohol photore-
channel for anatase-based nanosystems. The study of structural forming reaction: (i) the alcohol must have a hydrogen in the
defects closely related to morphology variables such as size, shape, ␣-position; (ii) primary alcohols undergo decarbonylation yield-
secondary particle size and porosity play a key role in control- ing a single molecule of CO2 and hydrogen plus an alkane (the
ling light-matter interactions. Hence, numerous studies have been exception being methanol which produces no alkane); (iii) methy-
devoted to overcome such problems, which include doping with lene groups produced in the reaction undergo complete oxidation
metal or non-metal elements [11–14], controlling the structure and to yield carbon dioxide; (iv) the dominant pathway for methyl
facets [15–17], and introducing defects into nanocrystals [18].
There are two main approaches for suppressing the reverse
reaction: either the use of sacrificial reagents or the creation of a
physical separation between the corresponding photoactive sites
on the surface of the photocatalyst. In this sense, the separation of
the photoactive sites would require surface separation of the photo-
generated electrons and holes which is commonly accomplished by
the use of specific co-catalyst. This point will be explicitly discussed
in Section 5.
As widely stated, water splitting reaction on TiO2 as model
semiconductor still shows a very low efficiency. Thermodynami-
cally, oxidation of water to produce O2 is more feasible than the
formation of peroxo and hydroxyl species. However, due to the
slow kinetics of the O2 evolution reaction, the formation of such
alternative species appears to become kinetically competitive with
O2 production. For this reason sacrificial agents such as methanol
or EDTA could be used to improve hydrogen productivity. Thus
the oxidation of sacrificial organic compounds by these highly
Fig. 4. Schematic model of H2 evolution reaction on core/shell noble-metal/Cr2 O3
reactive species would enhance the performance of the reduction
particulate system as a co-catalyst for photocatalytic overall water splitting
half-reaction. If we consider renewable compounds as sacrificial (Reprinted with permission from Ref. [94]. Copyright (2009) American Chemical
target molecules the environmental interest of the process would Society).
G. Colón / Applied Catalysis A: General 518 (2016) 48–59 51

Fig. 5. (A) (a–f) SEM images of CdS-Te heterostructures prepared at reaction times of (a) 6 min, (b) 6.25 min, (c) 6.5 min, (d) 6.75 min, (e) 7.0 min, and (f) 10.0 min. (g) Growth
scheme for the shish-kebab-like CdS-Te heterostructures; (B) Comparison of the photocatalytic activities of (a) CdS, a CdS/Te mixture, and CdS-Te heterostructures and (b)
CdS-Te, NiS/CdS-Te, MoS2 /CdS-Te, and CdS-Te@Pt,Pd in a lactic acid aqueous solution under visible-light illumination (Reprinted with permission from Ref. [102]. Copyright
(2015) American Chemical Society).

Fig. 6. (A) Kubelka–Munk absorption spectra, photographs (the left inset) and Tauc plots (the right inset) of the as-synthesized samples D52, D55, D60 and D64.(b); (B) H2
evolution from samples (30 mg) in methanol (20 vol%) aqueous solution at pH 4.5 under visible light (>400 nm) irradiation. Samples were platinated with 1% mass Pt prior to
irradiation. Inset: relationship between the reaction rate and the calcination temperature (Reprinted from Ref. [115] with permission from the Royal Society of Chemistry.

groups produced at the surface from the photo-reforming process 3. Photoreforming from oxygenates
is recombination with hydrogen and desorption. Other operational
parameters such as alcohol concentration and incident light inten- As above pointed out, the production of hydrogen by a photore-
sity have been reported to condition the progress of the reaction forming process shows two clear advantages with respect to water
[30,31]. splitting reaction. Firstly the use of sacrificial agent which could
In spite of the huge amount of results concerning to the hydro- help in the reduction half-reaction and secondly the use of biomass
gen production by photoreforming, the productivity obtained in derived substrates would provide an added-value to the overall
each case is quite difficult to compare due to the high variety process [32]. This coupled process represents an environmental
of operational conditions (reactor type, flows, lamp types and friendly and cost effective method for wastewater treatment, with
power,. . .). This fact would point out the strong necessity to stan- simultaneous production of energy.
dardize the results by providing the photonic efficiency.
52 G. Colón / Applied Catalysis A: General 518 (2016) 48–59

During photoreforming process, instead of water the organic was two times higher than that over Pt/TiO2 and the reaction selec-
compounds can act as scavengers of photoinduced holes and photo- tively lasted for a long time without deactivation [45].
generated oxygen, thereby retarding electron-hole recombination
as well as hydrogen–oxygen backward reaction. This way the
resulting hydrogen production rates become enhanced. As it can 4. Different H2 producing photocatalysts
be expected, in this case when complete oxidation of the sacrificial
agent is achieved, the rate of hydrogen production would drop to Since the first works describing the photocatalytic water
values comparable to those obtained from pure water. splitting reaction the number of systems proposed has been expo-
Within this framework, the use of biomass derived compounds nentially grown [46,47]. As mentioned in the previous section,
such as carbohydrates is gaining increasing interest as an abundant, suitable materials should fit the band energy conditions in order to
renewable and clean feedstock for hydrogen production [33,34]. proceed the water oxidation and reduction reactions. Considering
Kawai et al. firstly demonstrated the possibility to convert carbo- these electronic pre-requirements many photoactive semiconduc-
hydrates, not only sugar or starch but also cellulose, in the presence tors have been studied (Fig. 1).
of water and a RuO2 /TiO2 /Pt into hydrogen [35]. From this star- Among these TiO2 , SrTiO3 , and La:NaTaO3 were firstly studied
ing result many attempts have been reported considering different though their major disadvantage is that they only absorb light in
target molecules and catalysts. In this sense, as a general trend, the ultraviolet region of the solar spectrum [48,49]. However, in
the complex molecular structure of carbohydrates with respect to spite of the large number of the proposed systems, only a few actu-
simple alcohols (methanol or ethanol) leads to a lower productiv- ally show an appreciable activity for water photolysis. Recently,
ity. For this reason the mechanism of carbohydrate full degradation other UV-active systems such as Nb- and Ta-based oxides have
still need of further studies in order to be improved. been also considered for this reaction [50]. In this sense, in a recent
Thus, Fu et al. investigated the photocatalytic reforming of glu- paper, Soldat et al. proposed new types of BaTaO3 system prepared
cose over different noble metal supported TiO2 . They concluded by a citrate route. The obtained photocatalysts showed interest-
that glucose molecule would be bonded to the surface undercood- ing hydrogen production rates associated to the improved charge
inated Ti atoms through its hydroxyl groups. Once adsorbed, the carrier separation [51].
glucose molecule would trap a photogenerated hole, oxidizing and In the last years most relevant results regarding such systems
forming a RCH2 O• radical. This radical would continue the reac- are related to TiO2 based photocatalysts. In this case the research
tion through a further attack over another glucose molecule and has been focused in the structural and morphological modification
by a subsequent deprotonation and oxidation carboxylate species of TiO2 , the development of two or more components photocatalyst
could be formed. Finally, the [R COOH]− decarboxylates via a photo- that are more complicated than TiO2 or its the combination with a
Kolbe reaction resulting in eventual CO2 . On the other hand, the H+ wide range of possible co-catalysts.
from glucose deprotonation will be transferred to the Pt particles Thus, it has been reported that the particular morphology of TiO2
and then reduced to H2 by photogenerated electrons. They also could have a notable effect on the H2 production (Fig. 2) [52].
stated that the rate of hydrogen production is strongly dependent Teng et al. argued that owing to the higher surface energy of the
on the initial pH of the solution, reaching the maximum produc- {0 0 1} TiO2 facets, nanosheet-based photocatalyst shows higher
tion at higher pH values (pH > 11). Glucose shows a pKa value of ca hydrogen production activity and recyclability than the nanorod-
12.3 so when the pH is far below this point, glucose is present in its based one.
molecular form bonding to TiO2 surface by means of the hydroxyl In the same direction, Wang et al. also pointed out that hierarchi-
groups. At higher pH the molecule dissociates and the anionic form cal anatase TiO2 nanoflowers consisting of {0 0 1} facet dominating
of the molecule is able to capture more efficiently the photogener- nanosheets shows improved photocatalytic H2 evolution [53]. The
ated holes. At pH values higher than the pKa glucose molecule turns coexistence of different exposed planes could promote electron
negatively charged as TiO2 surface hampering the adsorption and transfer from {0 0 1} to {0 1 0} when they are photoexcited enhanc-
therefore the H2 evolution rate. Furthermore, the existence of ␣ and ing the photoexcited carrier lifetimes which is closely linked to the
␤ anomericity for saccharide molecule has also certain influence in number of {0 0 1}–{0 1 0} quasi-heterojunctions. Similarly, Wang
the reaction rate probably due to the way the molecule is adsorbed et al. also showed that the preparation of TiO2 particles with a high
at the surface [36]. Taking into account these mechanistic consid- percentage of reactive facets (∼85%) leads to 1.4–5 times higher
erations, as expected the hydrogen production by photoreforming performance than their counterparts in photocatalytic hydrogen
of mono-, di- and poly-saccharides decreased with increasing the production [54]. On the contrary, Gordon et al. proposed exactly the
molecular complexity [37]. opposed trend [55]. These authors argued that the photocatalytic
Another possibility arises from the use of different oxygenated activities of oxygen deficient anatase Pt-TiO2 showing preferen-
substrates non derived in this case from biomass such as waste tially {1 0 1} facet exposed are more active for the production
pollutants or even methane. In the first case, the possibility of the of hydrogen from methanolic solutions under solar illumination.
assembly two interesting processes, hydrogen formation and pol- Therefore the control of the reactive facet of TiO2 has clearly turned
lutant degradation, appears of great relevance [38]. Thus, a wide as a promising strategy for increasing the hydrogen production that
variety of organic pollutants in water has been photoreformed needs further studies.
achieving in all cases to complete mineralization simultaneously Besides these specific studies concerning to the exposed facets,
to the production of H2 . These examples include the degradation the origin of the intrinsic photoactivity for hydrogen production
of dyes [39], aldehydes [40], phenolic compounds [41] and even related to the structural and/or morphological features of TiO2 is
olive mill wastewater [42]. As in the case of biomass derived sub- still not clear enough. In a recent paper, Montes-Navajas et al. tried
strates, as the molecule is simpler the efficiency of the H2 evolution propose a correlation between the structural and morphological
is higher. features of different bare commercial TiO2 (without any supported
Finally the methane photoreforming has been also proved as co-catalyst) [56]. Unfortunately, from their results it was possible
a feasible way to produce H2 [43,44]. When Pt/TiO2 was photoir- to conclude that upon solar irradiation the photocatalytic activity
radiated in the flow of water vapour and methane around room tendency appears difficult to rationalize. These authors found that
temperature, hydrogen and carbon dioxide were the main prod- it was not possible to correlate well with crystallographic phases
ucts, and only trace amounts of ethane and carbon monoxide were present and it seems to derive from more complex interplay of
also observed.43 The hydrogen production rate over Pt/NaTaO3 :La different factors such as the presence of impurities, surface area,
G. Colón / Applied Catalysis A: General 518 (2016) 48–59 53

[85], WO3 [86,87], CdS and g-C3 N4 [88,89] have been explored
as potential photocatalysts. In this sense, (Ga1−x Znx )(N1−x Ox ) sys-
tem has been extensively studied by Domen’s group for the overall
water splitting reaction [90,91]. This solid solution alone shows a
scarce photoactivity, however when it is loaded with a specific co-
catalysts the reaction proceeds with a high hydrogen yield [92]. By
the moment, this system loaded with Rh2−y Cry O3 co-catalyst shows
the best visible light photocatalyst for water splitting, reaching a
quantum efficiency of 5.6% under visible light irradiation (Fig. 4)
[93,94].
Among the new photocatalysts for hydrogen production
recently proposed in the literature, perhaps the most interesting
ones are metal sulphides and graphitic carbon nitride. This par-
ticular interest is based in their notably photoactivity under visible
light irradiation. The origin of the visible photoactivity of metal sul-
phides can be attributed to the participation of S 3p orbitals (being
more negative than O 2p) on the construction of the valence band,
which increase the width of the band itself. Therefore, in contrast to
Fig. 7. Proposed electronic diagram for TiO2 /g-C3 N4 heterostructures for the pho- metal oxides many of them (CdS, MoS, AgS) exhibit a narrow band
tocatalytic hydrogen production (Reprinted with permission from Ref. [72]). gap. In spite of their remarkable visible photoactivity, there exists
an inherent drawback, i.e., photocorrosion in which chalcogenide
degree of crystallinity, particle size and morphology. Thus, from all itself could be oxidized by the photogenerated holes [95]. Among
the TiO2 samples studied it results that Evonik P25 showed the best the available sulphides, CdS with band gap of 2.4 eV is probably the
H2 production. In this sense, Obregón et al. recently point out the best-studied metal sulphide photocatalyst for hydrogen produc-
strong influence of structural and morphological features of TiO2 tion reaction, although its high toxicity upon photocorrosion. Such
on the Cu active sites with the hydrogen evolution production [57]. problem greatly obstructs its practical application due firstly to the
In spite of the complexity in certain cases, heterostructured sys- loss of activity and more importantly to the metal leakage pollution
tems provide an attractive and flexible approach for enhancing the (particularly serious in the case of cadmium) produced during the
reaction efficiency. Two or more semiconductors coupling with the reaction. It is believed that photocorrosion occurs in these chalco-
appropriate electronic band positions would lead to a more effi- genide photocatalysts when their activity irreversibly decreases
cient separation of the photoinduced charge carriers [58–60]. As even with the re-addition of sacrificial agent. Various efforts have
discussed in the previous section, for the particular case of H2 pro- been made to improve the stability of sulphide-based photocata-
duction through proton reduction, the studied systems must have lysts and retain its high photoactivity. The use of specific sacrificial
conduction band levels sufficiently higher than the H+ /H2 redox reagent which would hinder the sulphide photocorrosion process
potential. Following this approach different heterojunctions based is widely reported as a first attempt to stabilize the sulphide struc-
on TiO2 coupled with other semiconductors such as CdS [61–63], ture against photocorrosion. Thus, S2− , SO4 2− or S2 O3 2− species
ZnO [64], SnO2 [65], WO3 [66], Bi2 O3 [67], Fe2 O3 [68], ZrO2 [69,70], have been extensively used as sacrificial hole scavenger [46]. In
MoO3 [71], g-C3 N4 [72,73] or graphene [74,75] to form heterojunc- the presence of such species, chalcogenides show a remarkable
tions. Among these proposed composite systems CdS decorated photoactivity for water reduction.
TiO2 have received considerable attention due to the appropriate Regarding to the photocatalytic systems, the proposed
band alignment which leads to a highly efficient electron trans- approaches passed through the coupling sulphides with wide-
fer. In this sense Fang et al. proposed a ternary design in which a bandgap semiconductor or oxidation co-catalyst [96], and
transfer path for the photoexcited electrons of CdS to the core Au constructing a solid solution [97–99], complex core–shell [100] or
particles via the TiO2 nanocrystal bridge effectively suppresses the shelter [101] heterostructures which could protect the unstable
electron-hole recombination on the CdS photocatalyst [62]. Sim- semiconductor inside the core. An exotic example of the prepara-
ilarly, Zhou et al. proposed different CdS/M/TiO2 (M = Au, Ag, Pt, tion of complex heterostructures is reported by Hu et al. [102] These
Pd) ternary heterojunctions for efficient photocatalytic hydrogen authors described a series of multi-heterostructured metal chalco-
evolution [76] genides (CdS-Te, NiS/CdS-Te, and MoS2 /CdS-Te) with a surprising
Furthermore the different band structure of TiO2 polymorphs shish-kebab-like morphology synthesized by a microwave-assisted
has been also used, in this case for homojunction formations, show- pyrolysis of dithiocarbamate precursors (Fig. 5). CdS-Te and CdS-
ing enhanced charge separation (Fig. 3). Thus, anatase-rutile and Te-based composites exhibited enhanced photocatalytic activity
anatase-brookite heterostructures has been reported to exhibit and photostability for water splitting into H2 .
better H2 production with respect to the single TiO2 catalysts In recent years, a considerable attention has been paid to CuInS2
[77–82]. due to its environmentally benign metal components with respect
Indeed, by choosing the adequate anatase-rutile fraction and to CdS based photocatalysts and excellent solar-harvesting prop-
modification of in the TiO2 it is possible not only to enhance the erties. CuInS2 is a direct bandgap (∼1.5 eV) semiconductor which
reaction rate but also supress the CO production during photore- has a large optical absorption coefficient (>105 cm−1 ) and is con-
forming reaction [83]. sidered a promising semiconductor for photovoltaic devices and
This selectivity improving could be caused by the modifica- other optoelectronic applications [103]. As in the case of CdS dif-
tion of the surface acidity/basicity features of the homojunction. ferent heterostructured systems complex have been reported in
Therefore the control of this simple homojunction would provide the last years in order to enhance its photocatalytic performance
interesting additional benefits to hydrogen production reaction. [104,105].
Besides traditional oxide photocatalysts mostly based on TiO2 As an example, a complex ternary solid solution formed by
systems different alternatives have been studied. Thus, for the sake AgInS2 -CuInS2 -ZnS was proposed by Tsuji et al . some years ago
to propose alternative materials able to use the visible region of the [106]. Then, other solid solutions based on CuInS2 have been
solar spectrum other systems like (Ga1−x Znx )(N1−x Ox ) [84], Fe2 O3 reported. Thus, Ren et al . described the preparation of a (CuIn)x
54 G. Colón / Applied Catalysis A: General 518 (2016) 48–59

Fig. 8. Schematic illustration of photocatalytic water splitting over a semiconductor photocatalyst loaded with H2 − and O2 − evolution co-catalysts. (Reprinted from Ref.
with permission from the Royal Society of Chemistry).

Cd2(1−x) S2 solid solution by a hydrothermal method [107]. Loading exhibiting an n-type character. That means that upon excitation,
of Pt as co-catalyst results in a substantial improvement in H2 evo- they would readily accumulate excited holes on their surfaces
lution (2456 ␮mol/h · g) with respect to Pt-loaded pure CdS (40.2 showing high oxygen evolution activity. In this case, the coun-
␮mol/h · g). Moreover, these authors reported that the activity of terpart excited electrons would prefer to stay in the bulk of the
this system shows a high stability in the photocatalytic process after semiconductor; as a result, metal nanoparticles can often effec-
long use experiments. tively transport such electrons to the surface by guiding them along
Considering the high number of research works publish in the the metal–semiconductor interface [116]. Thus, the participation
last years, the second promising candidate for hydrogen production of a co-catalyst has been extensively considered for the hydrogen
reaction is g-C3 N4 [108]. Thus a high variety of complex sys- production reaction which could improve either reduction or oxi-
tems have been proposed for the optimization of its photocatalytic dation half-reactions (Fig. 8) [117,118]. Although along this section
performance. These approaches passed through the formation of we will focus our attention to reduction co-catalyst due the great
surface coupling hybridization with other semiconductors (such as impact in the H2 production reaction, it can be highlighted that
TiO2 [109], SnO2 [110], CdS [111], or graphene [112]), construction for O2 evolution reaction several co-catalysts (such as MnOx, IrOx,
of mesoporous structures [113], doping with metal (Fe, Ag, Au, Pd) CoOx or CoPi) have been also proposed [72,117,119].
or non-metal species (S, B and P) or sensitizing with organic dyes As mentioned, practically the whole studied photocatalytic sys-
[114]. Thus, although the intrinsic photocatalytic performance of tems need a high-active co-catalyst for reducing the overpotential
g-C3 N4 itself can be considered low, the control of the synthetic of proton reduction and enhancing hydrogen evolution efficiency.
parameters as well as the designing of complex heterostructures For the wide number of metal co-catalysts discussed, different
leads in most of the cases to a notably photoactivity compared to roles have been proposed in the literature which tried to explain
the former components. the function of such co-catalysts in the reaction. The first role
In this sense, Wu et al. reported that the photocatalytic activity invoked concerns to the improvement of the electron-hole life-
of g-C3 N4 is strongly related to the presence of structural defects time by trapping of photogenerated travelling electrons. Another
generated during the calcination process. These defects produce possibility argued is related to its participation in the reaction
additional optical absorption in the visible spectra of g-C3 N4 and of the organic with the surface, usually through dehydrogena-
weakened photoluminescence [115]. These authors demonstrated tion/decarbonylation reactions [120]. Even more, Joo et al. recently
that the structural control could have a strong impact on the final proposed an alternative striking pathway that does not involve
hydrogen production of g-C3 N4 (Fig. 6). electron transfer to the metal but requires the metal co-catalyst to
The photoactivity enhancement could also be achieved by the act as a catalyst for the recombination of the hydrogen atoms made
formation of heterostructure systems. By coupling process, g-C3 N4 via the reduction of protons on the surface of the semiconductor
based heterostructures can be formed by combining with visible instead [121].
light active semiconductor materials (such as CdS, Bi2 WO3 , and Till now this task assumed for reduction co-catalysts has been
BiOI), leading to a more efficient photogenerated charge separation reasonably accomplished by noble metals (Pt, Au, Ag or Pd). The
if the band diagram of both fit adequately. most accepted mechanism explaining the interaction between the
In addition, it can combine with large band gap photocatalysts metal and the active semiconductor involves the creation of a
(such as TiO2 , ZnO, and ZnWO4 ), which can largely broaden the Schottky barrier that would inject electrons from the semiconduc-
application of the g-C3 N4 based nanocomposites (Fig. 7). tor conduction band toward the metal Fermi level [122]. By driving
the photogenerated electrons to the metal active sites they serve
as cathodic centers for H2 production. This leads to the increase
5. The co-catalyst factor
in the electron transfer rate to the adsorbed species, retarding the
possibility of their recombination with holes. Noble metals gen-
As discussed in Section 2, in order to achieve a large-scale and
erally show a more stable metallic state which would prevent its
sustainable hydrogen production process from water, the photo-
oxidation during water splitting reaction making them useful as
catalytic system usually requires a high efficient photocatalyst.
active co-catalysts. The main drawback for a widespread use is
Therefore, most of the proposed photocatalyst are oxide type
G. Colón / Applied Catalysis A: General 518 (2016) 48–59 55

their scarcity and the subsequent high price of them. For this rea- traditional method. In this sense, Schweinberger et al. proposed
son, alternative transition metal co-catalysts such as Cu, Ni or Fe a successful method for preparing Pt clusters smaller than 2 nm
have been widely proposed [123]. However, metals in an oxidised supported on CdS surface by means of laser ablation [134]. These
state are expected to follow a different electron transfer mechanism authors found that the maximum hydrogen production rate was
due to the loss of metallic properties. Moreover, in recent papers achieved by Pt46 cluster and described a clear size effect within
it has been proposed transition sulphides such as CuS [124], NiS the studied sizes. Moreover, such size effect has been interpreted
[125] or MoS [126] as effective hydrogen production co-catalysts. by considering two electron transfer steps: first from the semicon-
However, as it was stated previously such candidates suffer dras- ductor to the cluster and second from the cluster to the hydrogen
tic anodic photocorrosion processes and trend to form S or sulfate atom.
[127]. Finally, the last approach to novel co-catalyst systems con- In a recent paper, Li et al. incorporated the metal particle local-
sist on bimetallic formulations by mixing different noble metals ization on the support as a crucial parameter to be considered [135].
or noble-transition metals. In the following sub-sections we will Thus, when PtO clusters co-catalyst is located on the (0 0 1) face of
describe in more details the results obtained by noble metal and TiO2 the H2 production rate observed was higher. This fact has been
transition metal based co-catalysts. related to shorter migration path of the photogenerated electrons
compared with Pt at (0 0 1) face. This would result in a rapid capture
5.1. Noble metal based co-catalysts of photogenerated electrons by metallic clusters and their avoiding
electron/hole recombination.
As it was mentioned before, the occurrence of dispersed metal In the case of gold, it is clear that deposition method affects in a
particles on the semiconductor catalyst would increase the elec- greater extent to the morphological features of the metal deposits
tron transfer to the adsorbed species and therefore enhance the [136]. It is widely reported that gold deposition method difficulty
efficiency of the process. This point is of great relevance for leads to small size and narrow distribution gold particles, espe-
the photocatalytic hydrogen production reaction. Historically, the cially when compared to platinum [137]. As mentioned before, the
traditional metal co-catalysts used were based on noble metal par- greater ability of Pt with respect to Au to act as electron sink and
ticles [128]. On the basis of the large literature describing the would be linked to the difference between their respective work
parameters that would govern the co-catalysts efficiency, a clear functions. This effect may be also associated to the smaller size of Pt
view of their role is still limited. nanoparticles with respect to Au ones. Furtherly, in addition to the
As pointed out by Yoshida et al. the efficiency of different noble particular metal morphological aspects the interface contact with
metal co-catalysts can be related to the work function of each one the support would be also of great importance. Thus, Jovic et al.
[32]. In principle, the work function of each metal should then play described that the creation of highly active three phase interfacial
a role in the formation of Schottky barrier. Taking into account sites “hot-spots” formed by the intimate contact of Au nanoparti-
this consideration, Pt appears as the most active noble metal for cles with anatase and rutile would lead to enhanced photoactivity
hydrogen generation. In spite of this general conclusion, the final [138].
hydrogen production efficiency is strongly dependent not only on A further aspect to be considered for gold and also silver
the type of metal co-catalyst but also on its nature and concen- based co-catalysts concerns the possibility of visible photoactivity
tration [129]. In this sense, the preparation procedure used for the through their plasmon excitation (Fig. 9) [139,140].
metal decoration could have great influence on its final photocat- It has been reported that the localized surface plasmon (LSP) has
alytic performance. In fact, even when we use the same metal, the important applications in photocatalysis. This is mainly due to the
deposition method clearly affects to the photoactivity [130,131]. following aspects: (i) direct injection of energetic electrons to semi-
As a general trend, the specific morphology (dispersion and metal conductors, resembling dye sensitizations [141]; (ii) enhancement
particle size) and oxidation state appears as the key parameters of electron–hole pairs generation in the regions of semiconductor in
which would condition the photocatalytic activity. Thus, Navarro vicinity to the metal nanoparticles, due to amplified near-field elec-
et al. stated that for particles in the metallic oxidation state the tromagnetic field and resonant photon scattering, i.e. via resonant
morphology does not have a relevant influence indicating that the energy transfer (RET) [142].
active site would be located in the metal-semiconductor inter- Within this excitation mode, the electronic mechanism and
face [131]. On the contrary, for partial oxidised metal particles it reaction scheme drastically change (Fig. 9). Upon visible light irradi-
has been observed improved photoactivities probably due to bet- ation TiO2 is totally inactive and the Au NPs would act preferentially
ter e− conductivity of the partially charged Pt particles that may as light harvesters injecting hot electrons into the conduction band
enhance the trapping electrons and thereby extending the elec- of TiO2 promoting water reduction to H2 . Considering this role
tron/hole carriers separation. In a recent paper Luo et al. stablished of light harvester, high visible light absorption by gold might be
an interesting correlation between electrocatalysis and photocatal- the control parameter for the overall photocatalytic efficiency. In
ysis by noble metal based materials [132]. Thus, since in electrodes this case, contrary to UV-irradiation conditions, the Au particle
for electrocatalytic applications the size and shape of noble metal size would not play a key role in the photocatalytic activity [143].
arise as a crucial parameter to be controlled, they proposed the This reverse electronic mechanism would imply that the reduction
same condition for photocatalytic materials. These authors argued site would be placed at TiO2 and the oxidation site on the metal
that conventional deposition methods (photodeposition or impreg- nanoparticle.
nation methods) do not allow controlling such morphological Recently, Kowalska et al. reported the action spectra for different
parameters. Thus, they obtained nano-cubic and nano-spherical Pt Au-TiO2 systems with varying gold and TiO2 sizes for the iso-
particles prepared via a shape-and-size-controlled synthesis and propanol oxidation reaction [144]. They demonstrated that under
loaded onto a CdS semiconductor. They finally stated that for both plasmonic excitation a direct relationship between the gold size
morphologies cluster size is the determining parameter, being the with the photocatalytic activity can be established. Thus, under
shape effect more significant for larger Pt particles than for smaller visible irradiation, the greater the gold particle size (e.g., broader
particles. In the same line, Xing et al. have synthesized isolated particle size distribution with larger mean and higher moments,
metal atoms (Pt, Pd, Rh, or Ru) stably by anchoring on TiO2 by a facile particularly the variance), the higher the efficiency exhibited due
one-step method [133]. Such isolated metal atom based photocat- to the larger number of photons absorbed. As above pointed out,
alysts show excellent stability for H2 evolution being 6–13-fold the higher performance observed in this case would be related
more photoactive than the metal clusters loaded on TiO2 by the to the extended plasmonic effect observed on large particles. On
56 G. Colón / Applied Catalysis A: General 518 (2016) 48–59

Fig. 9. Different excitation scheme for Au-TiO2 systems.

the contrary, under UV conditions, smaller gold particles clearly the photocatalytic hydrogen production and the textural character-
produce higher photoactivities. Very recently, suggested that the istics of a Cu-doped TiO2 system [157]. In this case the photoactivity
co-existence of large (>10 nm) and small (<5 nm) Au nanoparti- was correlated to the Cu nanostructure obtained and especially
cles in Au-TiO2 catalysis prepared by multistep photodeposition with deterioration of the electronic interactions between the com-
method contributed to stronger SPR photoabsorption, enhanced H2 ponents of metal − semiconductor composites with growth of the
evolution rate from 2-propanol in aqueous suspensions under vis- metal nanoparticles size. On this basis, the interaction with TiO2
ible light irradiation. In these cases, smaller Au nanoparticles acted is expected to stabilize the copper oxides against photocorrosion
as a reduction site for H2 evolution. and could enhance the activity by a most efficient process. Thus,
Similarly, silver showing a slightly lower work function than well dispersed CuOx nanoparticles would be easily reduced during
Pt or Au has been used as a suitable co-catalyst for H2 produc- the reaction forming well dispersed copper metal particles which
tion reaction.32 However, probably due to the lower photoactivities would favour electron trapping. Recently, Valero et al. prepared
obtained the number of published papers is scarce. two set of Cu-TiO2 systems by means of different deposition meth-
ods [158]. They proposed that easily reduced Cu2+ species obtained
preferentially by chemical reduction method could be the respon-
5.2. Non-noble metal co-catalysts sible of the higher photoactivity of Cu-TiO2 (Fig. 10).
Moreover, the final metal cluster size, dispersion and oxidation
As it has been discussed in the previous section, platinum has state could be strongly influenced by the structural and surface sup-
been frequently used onto semiconductor surfaces as a co-catalyst. port features [57]. Although Ni-TiO2 system has been also proven
However, if we pursue a scaling up production for water splitting- to be an alternative candidate to noble metal co-catalyst systems
hydrogen generation, the use of any platinum-based (or in general
noble metal based) material is quite questionable due to the scarcity
of this noble metal on the Earth’s crust which lead to an important
economic problem. In this sense recent important advances have
been achieved in the direction of mimicking the catalytic active
sites of natural hydrogenases which are based on earth-abundant
elements such as Cu, Fe, Co, Ni [145–150]. Among these transition
metal co-catalysts and although the reported rates till now are still
far from those obtained for noble metals, copper based catalysts
appear as a promising candidate [151]. In spite of the great num-
ber of papers dedicated to the study of Cu-based photocatalysts for
hydrogen generation, the nature of the active copper species (Cu2 O
or CuO) and other external factors governing its photoactivity is less
understood [152]. As already mentioned, the oxidised state of tran-
sition metal co-catalysts makes a difference with respect to noble
metal ones and includes an additional parameter to be considered.
Thus, within the discussion about the copper active species, some
authors claimed that Cu2 O species are responsible for the photoac-
tivity for H2 generation from water reduction [153]. Thus, Xi et al.
correlated the presence of easily reducible Cu2 O species toward
metallic Cu during the H2 production reaction with the improved
photocatalytic activity. Similarly, Jung et al. associated the higher
photoactivity of Cu-TiO2 systems to the dispersion and reducibility
of CuOx (Cu+ /Cu0 ) species [154]. On the other hand, other authors
argued that the presence of CuO would be responsible for the
enhanced separation of photoinduced electrons and holes [155].
In the same direction, Ampelli et al. reported the improving effect
of Cu2+ incorporation into the TiO2 structure and the simultaneous
creation of oxygen vacancies [147]. Thus, embedded CuOx@TiO2
Fig. 10. (A) H2 production for (a) CuSTi and (b) CuSTr series. (c) Calculated photonic
systems showed higher photocatalytic performances with respect
efficiencies for Cu-doped TiO2 ; (B) Schematic insight of copper situation on different
to impregnated samples [156]. Within this context, Korzhak et al. samples. (Reprinted with permission from Ref. [158]. Copyright (2014) American
investigated the relationships between the quantum efficiencies of Chemical Society).
G. Colón / Applied Catalysis A: General 518 (2016) 48–59 57

Fig. 11. (a) Photocatalytic H2 evolution using the as-prepared metal NPs supported on TiO2 catalyst materials. (b) The derived H2 production rates and apparent quantum
efficiencies (AQE) of the metal NPs supported on TiO2 catalyst materials (Reprinted from Ref. [164] with permission from the Royal Society of Chemistry).

little investigation in relation to photocatalytic H2 production can and size of co-catalyst metal particles have considerable influence
be found in the literature. As in the case of copper based systems, on the photocatalytic activity. These points would suggest that in
confusion exists as to whether NiO or Ni is the active co-catalyst addition to the electrocatalytic activity of surface redox reactions,
species [159,160] In a recent paper, Chen et al. presented interesting electronic interaction at the co-catalyst-semiconductor interface is
experimental evidences suggesting that highly dispersed metallic crucial.
Ni (∼1 nm sized clusters) would be the active co-catalyst species
[161]. Moreover, the 0.5 wt.% Ni/TiO2 photocatalyst showed supe- 6. Conclusion
rior activity to a 2 wt.% Au/TiO2 reference photocatalyst.
From the huge number of contributions published in the last
5.3. Bimetallic co-catalysts years it is clear that the photocatalytic hydrogen production reac-
tion from water is an attractive reaction. The potential conversion
As afore mentioned, in a metal-semiconductor configuration of solar energy and renewable resources (such as alcohols or
the excited electrons need overcome the Schottky in order to be biomass derivatives) into valuable fuels by photoreforming pro-
transferred from the semiconductor to the deposited metal. Thus cess is undoubtedly not only attractive but also very necessary
the height of the Schottky barrier would determine the efficiency under the point of view of the global sustainability. However,
of electron transfer. The extent of the Schottky barrier is directly much heavier efforts are needed to understand the behaviour of
related to the work function of metal as well as the conduction complex molecules such as carbohydrates. As it can be envisaged
band electron affinity of the semiconductor. Taking into consid- along this short overview the development of highly efficient pho-
eration this premise the work function of the metal could be tocatalytic material is by the moment a crucial task that might
tuned by the introduction of another metal with different work be strongly reinforced. On one hand the investigation of highly
function [162]. The formation of a bimetallic co-catalyst has been active semiconductors or heterostructured systems that could (i)
proposed to promote the photocatalytic activity of the TiO2 -based use as much as possible the solar radiation and (ii) optimize the
photocatalytic system instead of their monometallic counterparts. electronic mechanism of photogenerated pairs separation. Sec-
Several examples of this approach have recently appeared in the ondly the incorporation of efficient reduction (and even oxidation
literature considering noble metal and transition metal homo- co-catalysts, only slightly mentioned in this review) co-catalysts
and heterojunction. As monometallic co-catalyst platinum has economically suitable for this application that could enhance the
been extensively used since it shows the higher photocatalytic photocatalytic efficiency by improving the charge transfer to adsor-
performance for hydrogen production. However different bimetal- bates. By now, the low quantum efficiencies reported and the
lic formulations have been proposed denoting the wide field to absence of adequate scaling-up studies, lead to practical lack of
be explored. Thus, bimetallic Pt-Au [163,164] or Pt-Ag [165,166] industrial application. Unfortunately, we have to admit that Verne’s
have been recently proposed for the hydrogen production reaction prediction of a “solar fuels” industry from water is still no more
showing improved photoactivities with respect to the monometal- than a futuristic concept. For this reason the development of a suit-
lic co-catalyst (Fig. 11). able photocatalyst is extremely decisive and might encourage us to
It is clear that noble metals are expensive and rare candidates emphasize our efforts for this scope.
to be used even in low loading percentage, making their practi-
cal use certainly doubtful. The development of alternative noble Acknowledgments
metals-free substitutes have been discussed in the previous section
pointing out copper and nickel as promising alternative to Pt. The financial support by CTQ2014-60524-R project is fully
Upon this consideration, it can be deduced that the Cu–Ni acknowledged.
bimetallic co-catalyst might have an expected better performance
due to the synergy between both metals [167]. The study of these References
heterostructured co-catalysts in different configurations (core-
[1] A. Fujishima, K. Honda, Nature 238 (1972) 37–38.
shell, alloys) is opening new possibilities in the development of [2] J.D. Holladay, J. Hu, D.L. King, Y. Wang, Catal. Today 139 (2009) 244–260.
highly efficient photocatalysts for hydrogen generation. [3] T.L. LeValley, A.R. Richard, M. Fan, Int. J. Hydrogen Energ. 39 (2014)
On the basis of all the above discussion reported in the litera- 16983–17000.
[4] T. Hisatomi, K. Kubota, K. Domen, Chem. Soc. Rev. 43 (2014) 7520–7535.
ture it is clear that different parameters could strongly determine [5] P. Ruban, K. Sellappa, Energy 73 (2014) 926–932.
the role of co-catalyst in the photocatalytic hydrogen evolution [6] C.A. Rodríguez, M.A. Modestino, D. Psaltis, C. Moser, Energy Environ. Sci. 7
reaction. Thus, factors such as the loading amount, the dispersion (2014) 3828–3835.
58 G. Colón / Applied Catalysis A: General 518 (2016) 48–59

[7] B.A. Pinaud, J.D. Benck, L.C. Seitz, A.J. Forman, Z. Chen, T.G. Deutsch, B.D. [61] L. Qi, J. Yu, M. Jaroniec, Phys. Chem. Chem. Phys. 13 (2011) 8915–8923.
James, K.N. Baum, G.N. Baum, S. Ardo, H. Wang, E. Miller, T.F. Jaramillo, [62] J. Fang, L. Xu, Z. Zhang, Y. Yuan, S. Cao, Z. Wang, L. Yin, Y. Liao, C. Xue, ACS
Energy Environ. Sci. 6 (2013) 1983–2002. Appl. Mater. Interfaces 5 (2013) 8088–8092.
[8] T. Hisatomi, K. Takanabe, K. Domen, Catal. Lett. 145 (2015) 95–108. [63] J. Li, S.K. Cushing, P. Zheng, T. Senty, F. Meng, A.D. Bristow, A. Manivannan,
[9] A. Kubacka, M. Fernández-García, G. Colón, Chem. Rev. 112 (2012) N. Wu, J. Amer. Chem. Soc. 136 (2014) 8438–8449.
1555–1614. [64] A.M. Hussein, R.V. Shende, Int. J. Hydrogen Energ. 39 (2014) 5557–5568.
[10] A.L. Linsebigler, G. Lu, J.T. Yates, Chem. Rev. 95 (1995) 735–758. [65] R. Sasikala, A. Shirole, V. Sudarsan, T. Sakuntala, C. Sudakar, R. Naik, S.R.
[11] C. Chen, X. Burda, J. Am. Chem. Soc. 130 (2008) 5018–5019. Bharadwaj, Int. J. Hydrogen Energ. 34 (2009) 3621–3630.
[12] A.M. Maı́rquez, J.J. Plata, Y. Ortega, J. Fernández-Sanz, G. Coloı́n, A. Kubacka, [66] E. Karácsonyi, L. Baia, A. Dombi, V. Danciu, K. Mogyorósi, L.C. Pop, G. Kovács,
M. Fernaı́ndez-García, J. Phys. Chem. C 116 (2012) 18759–18767. V. Coşoveanu, A. Vulpoi, S. Simon, Z. Pap, Catal. Today 208 (2013) 19–27.
[13] A. Wang, H. Jing, Dalton Trans. 43 (2014) 1011–1018. [67] B. Naik, S. Martha, K.M. Parida, Int. J. Hydrogen Energ. 36 (2011) 2794–2802.
[14] A. Kubacka, G. Colón, M. Fernández-García, Catal. Today 143 (2009) 286–292. [68] I.A. Rodionov, E.V. Mechtaeva, I.A. Zvereva, Russ. J. Gen. Chem. 84 (2014)
[15] G. Tian, Y. Chen, W. Zhou, K. Pan, C. Tian, X.R. Huang, H. Fu, CrystEngComm. 611–616.
13 (2011) 2994–3000. [69] S. Onsuratoom, S. Chavadej, T. Sreethawong, Int. J. Hydrogen Energ. 36
[16] W.J. Ong, L.L. Tan, S.P. Chai, S.T. Yong, A.R. Mohamed, ChemSusChem. 7 (2011) 5246–5261.
(2014) 690–719. [70] Abdulmenan M. Hussein, Rajesh V. Shende, Int. J. Hydrogen Energ. 39 (2014)
[17] G. Liu, J.C. Yu, G.Q. Lu, H.M. Cheng, Chem. Commun. 47 (2011) 6763–6783. 5557–5568.
[18] M. Kong, Y. Li, X. Chen, T. Tian, P. Fang, X. Zheng, F. Zhao, J. Am. Chem. Soc. [71] B.J. Ma, J.S. Kim, C. Hyuck Choi, S.I. Woo, Int. J. Hydrogen Energ. 38 (2013)
133 (2011) 16414–16417. 3582–3587.
[19] X. Fu, J. Long, X. Wang, D.Y.C. Leung, Z. Ding, L. Wu, Z. Zhang, Z. Li, X. Fu, Int. [72] S. Obregón, G. Colón, Appl. Catal. B: Environ. 144 (2014) 775–782.
J. Hydrogen Energ. 33 (2008) 6484–6491. [73] J. Wang, J. Huang, H. Xie, A. Qu, Int. J. Hydrogen Energy 39 (2014) 6354–6363.
[20] C. Wang, J. Rabani, D.W. Bahnemann, J.K. Dohrmann, J. Photochem. [74] W. Fan, Q. Lai, Q. Zhang, Y. Wang, J. Phys. Chem. C 115 (2011) 10694–10701.
Photobiol. A: Chem. 148 (2002) 169–176. [75] Y. Yang, E. Liu, H. Dai, L. Kang, H. Wu, J. Fan, X. Hu, H. Liu, Int. J. Hydrogen
[21] A.V. Puga, A. Forneli, H. García, A. Corma, Adv. Func. Mater. 24 (2014) Energy 39 (2014) 7664–7671.
241–248. [76] H. Zhou, J. Pan, L. Ding, Y. Tang, J. Ding, Q. Guo, T. Fan, D. Zhang, Int. J.
[22] C. Ampelli, R. Passalacqua, C. Genovese, S. Perathoner, G. Centi, T. Montini, V. Hydrogen Energy 39 (2014) 16293–16301.
Gombac, J.J. Delgado Jaen, P. Fornasiero, RSC Adv. 3 (2013) 21776–21788. [77] O. Rosseler, M.V. Shankar, M.K. Du, L. Schmidlin, N. Keller, V. Keller, J. Catal.
[23] V.M. Daskalaki, D.I. Kondarides, Catal. Today 144 (2009) 75–80. 269 (2010) 179–190.
[24] V.M. Daskalaki, P. Panagiotopoulou, D.I. Kondarides, Chem. Engineer. J. 170 [78] K. Connelly, A.K. Wahab, H. Idriss, Mater. Renew. Sustain. Energy 1 (2012) 3.
(2011) 433–440. [79] F. Cai, Y. Tang, H. Shen, C. Wang, A. Ren, L. Xiao, W. Gu, W. Shi,
[25] A. Speltini, M. Sturini, D. Dondi, E. Annovazzi, F. Maraschi, V. Caratto, A. CrystEngComm 17 (2015) 1086–1091.
Profumo, A. Buttafava, Photochem. Photobiol. Sci. 13 (2014) 1410–1419. [80] Q. Tay, X. Liu, Y. Tang, Z. Jiang, T.C. Sum, Z. Chen, J. Phys. Chem. C 117 (2013)
[26] L. Zhang, J. Shi, M. Liu, D. Jing, L. Guo, Chem. Comm. 50 (2014) 192–194. 14973–14982.
[27] N. Luo, Z. Jiang, H. Shi, F. Cao, T. Xiao, P.P. Edwards, Int. J. Hydrogen Energy [81] J. Cihlar, V. Kasparek, M. Kralova, K. Castkova, Int. J. Hydrogen Energy 40
34 (2009) 125–129. (2015) 2950–2962.
[28] L.F. Wei, X.J. Zheng, Z.H. Zhang, Y.J. Wei, B. Xie, M.B. Wei, X.L. Sun, Int. J. [82] V. Gombac, L. Sordelli, T. Montini, J.J. Delgado, A. Adamski, G. Adami, M.
Energy Res. 36 (2012) 75–86. Cargnello, S. Bernai, P. Fornasiero, J. Phys. Chem. A 114 (2010) 3916–3925.
[29] H. Bahruji, M. Bowker, P.R. Davies, F. Pedrono, Appl. Catal. B: Environ. 107 [83] Q. Xu, Y. Ma, J. Zhang, X. Wang, Z. Feng, C. Li, J. Catal. 278 (2011) 329–335.
(2011) 205–209. [84] K. Maeda, T. Teramura, M. Takata, N. Hara, K. Saito, Y. Toda, H. Inoue, K.
[30] G.N. Nomikos, P. Panagiotopoulou, D.I. Kondarides, X.E. Verykios, Appl. Kobayashi, J. Phys. Chem. B 109 (2005) 20504–20510.
Catal. B: Environ. 146 (2014) 249–257. [85] V. Preethi, S. Kanmani, Int. J. Hydrogen Energy 39 (2014) 1613–1622.
[31] G.L. Chiarello, D. Ferri, E. Selli, J. Catalysis 280 (2011) 168–177. [86] A.B.D. Nandiyanto, O. Arutanti, T. Ogi, F. Iskandar, T.O. Kim, K. Okuyama,
[32] K. Shimura, H. Yoshida, Energy Environ. Sci. 4 (2011) 2467–2481. Chem. Eng. Sci. 101 (2013) 523–532.
[33] D.I. Kondarides, V.M. Daskalaki, A. Patsoura, X.E. Verykios, Catal. Lett. 122 [87] H. Katsumata, Y. Tachi, T. Suzuki, S. Kaneco, RSC Adv. 4 (2014) 21405–21409.
(2008) 26–32. [88] X.L. Wang, W.Q. Fang, Y. Yao, P. Liu, Y. Wang, H. Zhang, H. Zhao, H.G. Yang,
[34] A. Gallo, T. Montini, M. Marelli, A. Minguzzi, V. Gombac, R. Psaro, P. RSC Adv. 5 (2015) 21430–21433.
Fornasiero, V. Dal Santo, ChemSusChem. 5 (2012) 1800–1811. [89] X. Li, A.F. Masters, T. Maschmeyer, ChemCatChem. 7 (2015) 121–126.
[35] T. Kawai, T. Sakata, Nature 286 (1980) 474–476. [90] K. Maeda, H. Hashiguchi, H. Masuda, R. Abe, K. Domen, J. Phys. Chem. C 112
[36] M. Zhou, Y. Li, S. Peng, G. Lu, S. Li, Catal. Comm. 18 (2012) 21. (2008) 3447–3452.
[37] C.G. Silva, M.J. Sampaio, R.R.N. Marques, L.A. Ferreira, P.B. Tavares, A.M.T. [91] T. Ikeda, A. Xiong, T. Yoshinaga, K. Maeda, K. Domen, T. Teranishi, J. Phys.
Silva, J.L. Faria, Appl. Catal. B: Environ. 178 (2015) 82–90. Chem. C 117 (2013) 2467–2473.
[38] C. Liu, Z. Lei, Y. Yang, Z. Zhang, Water Res. 47 (2013) 4986–4992. [92] T. Takata, C. Pan, K. Domen, Sci. Technol. Adv. Mater. 16 (2015) 033506.
[39] J. Kim, Y. Park, H. Par, Int. J. Photoenergy (2014) 324859. [93] K. Maeda, K. Teramura, D.L. Lu, T. Takata, N. Saito, Y. Inoue, K. Domen, Nature
[40] A. Patsoura, D.I. Kondarides, X.E. Verykios, Catal. Today 124 (2007) 94–102. 440 (2006) 295.
[41] J. Kim, D. Monllor-Satoca, W. Choi, Energy Environ. Sci. 5 (2012) 7647–7656. [94] M. Yoshida, K. Takanabe, K. Maeda, A. Ishikawa, J. Kubota, Y. Sakata, Y.
[42] A. Speltini, M. Sturini, F. Maraschi, D. Dondi, G. Fisogni, E. Annovazzi, A. Ikezawa, K. Domen, J. Phys. Chem. C 113 (2009) 10151–10157.
Profumo, A. Buttafava, Int. J. Hydrogen Energy 40 (2015) 4303–4310. [95] M. Ashokkumar, Int. J. Hydrogen Energy 23 (1998) 427–438.
[43] H. Yoshida, K. Hirao, J. Nishimoto, K. Shimura, S. Kato, H. Itoh, T. Hattori, J. [96] Z. Yan, H. Wu, A. Han, X. Yu, P. Du, Int. J. Hydrogen Energy 39 (2014)
Phys. Chem. C 112 (2008) 5542–5551. 13353–13360.
[44] K. Shimura, H. Kawai, T. Yoshida, H. Yoshida, ACS Catal. 2 (2012) 2126–2134. [97] C.C. Chan, C.C. Chang, C.H. Hsu, Y.C. Weng, K.Y. Chen, H.H. Lin, W.C. Huang,
[45] K. Shimura, S. Kato, T. Yoshida, H. Itoh, T. Hattori, H. Yoshida, J. Phys. Chem. S.F. Cheng, Int. J. Hydrogen Energy 39 (2014) 1630–1639.
C 114 (2010) 3493–3503. [98] M. Liu, L. Zhang, X. He, B. Zhang, H. Song, S. Li, W. You, J. Mater. Chem. A 2
[46] Y. Kudo, A. Miseki, Chem. Soc. Rev. 38 (2009) 253–278. (2014) 4619–4626.
[47] I.E. Castelli, D.D. Landis, K.S. Thygesen, S. Dahl, I. Chorkendorff, K.W. [99] Q. Li, H. Meng, P. Zhou, Y. Zheng, J. Wang, J. Yu, J. Gong, ACS Catalysis 3
Jaramillo, T.F. Jacobsen, Energy Environ. Sci. 5 (2012) 9034–9043. (2013) 882–889.
[48] R.G. Carr, G.A. Somorjai, Nature 290 (1981) 576–577. [100] J. Zhang, Y. Wang, J. Jin, J. Zhang, Z. Lin, F. Huang, J. Yu, ACS Appl. Mater.
[49] H. Kato, A. Kudo, Chem. Phys. Lett. 295 (1998) 487–492. Interfaces 5 (2013) 10317–10324.
[50] A.A. Ismail, D.W. Bahnemann, Sol. Energ. Mat. Sol. Cells 128 (2014) 85–101. [101] Y. Tang, X. Hu, C. Liu, Phys. Chem. Chem. Phys. 16 (2014) 25321–25329.
[51] J. Soldat, R. Marschall, M. Wark, Chem. Sci. 5 (2014) 3746–3752. [102] J. Hu, A. Liu, H. Jin, D. Ma, D. Yin, P. Ling, S. Wang, Z. Lin, J. Wang, J. Am.
[52] F. Teng, M. Chen, N. Li, X. Hua, T. Wang, K. Xu, ChemCatChem. 6 (2014) Chem. Soc. 137 (2015) 11004–11010.
842–847. [103] L. Li, N. Coates, D. Moses, J. Am. Chem. Soc. 132 (2010) 22–23.
[53] D. Wang, P. Kanhere, M. Li, Q. Tay, Y. Tang, Y. Huang, T.C. Sum, N. Mathews, [104] X. Zhang, Y. Du, Z. Zhou, L. Guo, Int. J. Hydrogen Energy 35 (2010)
T. Sritharan, Z. Chen, J. Phys. Chem. C 117 (2013) 22894–22902. 3313–3321.
[54] B. Wang, X.Y. Lu, L.K. Yu, J. Xuan, M.K.H. Leung, H. Guo, CrystEngComm. 16 [105] X. Yu, X. An, A. Shavel, M. Ibáñez, A.J. Cabot, J. Mater. Chem. A 2 (2014)
(2014) 10046–10055. 12317–12322.
[55] T.R. Gordon, M. Cargnello, T. Paik, F. Mangolini, R.T. Weber, P. Fornasiero, [106] I. Tsuji, H. Kato, A. Kudo, Angew. Chem. Int. Ed. 44 (2005) 3565–3568.
C.B. Murray, J. Am. Chem. Soc. 134 (2012) 6751–6761. [107] L. Ren, F. Yang, Y.R. Deng, N.N. Yan, S. Huang, D. Lei, Q. Sun, Y. Yu, Int. J.
[56] P. Montes-Navajas, M. Serra, A. Corma, H. García, Catal. Today 225 (2014) Hydrogen Energy 35 (2010) 3297–3305.
52–54. [108] S. Cao, J. Yu, J. Phys. Chem. Lett. 5 (2014) 2101–2107.
[57] S. Obregón, M.J. Muñoz-Batista, M. Fernández-García, A. Kubacka, G. Colón, [109] Z. Jiang, D. Liu, D. Jiang, W. Wei, K. Qian, M. Chen, J. Xie, Dalton Trans. 43
Appl. Catal. B: Environ. 179 (2015) 468–478. (2014) 13792–13802.
[58] D.Y.C. Leung, X. Fu, C. Wang, M. Ni, M.K.H. Leung, X. Wang, X. Fu, [110] Y. Zang, L. Li, X. Li, R. Lin, G. Li, Chem. Engineer. J. 246 (2014) 277–286.
ChemSusChem. 3 (2010) 681–694. [111] J. Zhang, Y. Wang, J. Jin, J. Zhang, Z. Lin, F. Huang, J. Yu, ACS Appl. Mater.
[59] Y.P. Yuan, L.W. Ruan, J. Barber, S.C.J. Loo, C. Xue, Energy Environ. Sci. 7 Interfaces 5 (2013) 10317–10324.
(2014) 3934–3951. [112] Q. Xiang, J. Yu, M. Jaroniec, J. Phys. Chem. C 115 (2011) 7355–7363.
[60] R. Marschall, Adv. Funct. Mater. 24 (2014) 2421–2440. [113] X. Li, A.F. Masters, T. Maschmeyer, ChemCatChem. 7 (2015) 121–126.
G. Colón / Applied Catalysis A: General 518 (2016) 48–59 59

[114] Z. Zhao, Y. Sun, F. Dong, Nanoscale 7 (2015) 15–37. [142] S.K. Cushing, J.T. Li, F.K. Meng, T.R. Senty, S. Suri, M.J. Zhi, J. Am. Chem. Soc.
[115] Po Wu, J. Wang, J. Zhao, L. Guo, F.E. Osterloh, J. Mater. Chem.A 2 (2014) 134 (2012) 15033–15041.
20338–20344. [143] M. Serra, J. Albero, H. García, ChemPhysChem. 16 (2015) 1842–1845.
[116] K. Domen, A. Kudo, T. Onishi, N. Kosugi, H. Kuroda, J. Phys. Chem. 90 (1986) [144] E. Kowalska, R. Abe, B. Ohtani, Chem. Commun. 24 (2009) 1–243.
292–295. [145] E. Pulido Melián, M. Nereida Suárez, T. Jardiel, J.M. Doña Rodríguez, A.C.
[117] J. Yang, D. Wang, H. Han, C. Li, Acc. Chem. Res. 46 (2013) 1900–1909. Caballero, J. Araña, D.G. Calatayud, O. González Díaz, Appl. Catal. B: Environ.
[118] J. Yang, H. Yan, X. Zong, F. Wen, M. Liu, C. Li, Phil. Trans. R Soc. A 371 (2013) 152–153 (2014) 192–201.
20110430. [146] R. Dholam, N. Patel, M. Adami, A. Miotello, Int. J. Hydrogen Energy 34 (2009)
[119] R. Li, H. Han, F. Zhang, D. Wang, C. Li, Energy Environ. Sci. 7 (2014) 5337–5346.
1369–1376. [147] V. Artero, M. Chavarot-Kerlidou, M. Fontecave, Angew. Chem. Int. Ed. 50
[120] M. Bowker, H. Bahruji, J. Kennedy, W. Jones, G. Hartley, C. Morton, Catal. (2011) 7238–7266.
Lett. 145 (2015) 214–219. [148] C. Ampelli, R. Passalacqua, C. Genovese, S. Perathoner, G. Centi, T. Montini, V.
[121] J.B. Joo, R. Dillon, I. Lee, Y. Yin, C.J. Bardeen, F. Zaera, PNAS 111 (2014) Gombac, J.J. Delgado Jaen, P. Fornasiero, RSC Adv. 3 (2013) 21776–21788.
7942–7947. [149] S. Oros-Ruiz, R. Zanella, S.E. Collins, A. Hernández-Gordillo, R. Gómez, Catal.
[122] A.L. Linsebigler, G. Lu, J.T. Yates, Chem Rev. 95 (1995) 735–758. Commun. 47 (2014) 1–6.
[123] H. Bahruji, M. Bowker, P.R. Davies, J. Kennedy, D.J. Morgan, Int. J. Hydrogen [150] Z. Wang, Y. Liu, D.J. Martin, W. Wang, J. Tang, H. Huang, Phys. Chem. Chem.
Energy 40 (2015) 1465–1471. Phys. 15 (2013) 14956–14960.
[124] Q. Wang, N. An, Y. Bai, H. Hang, J. Li, X. Lu, Y. Liu, F. Wang, Z. Li, Z. Lei, Int. J. [151] L. Clarizia, D. Spasiano, I. Di Somma, R. Marotta, R. Andreozzi, D.D.
Hydrogen Energy 38 (2013) 10739–10745. Dionysiou, Int. J. Hydrogen Energy 39 (2014) 16812–16831.
[125] J. Ran, J. Zhang, J. Yu, S.Z. Qiao, ChemSusChem. 7 (2014) 3426–3434. [152] A. Kubacka, M.J. Muñoz-Batista, M. Fernández-García, S. Obregón, G. Colón,
[126] J. Wang, B. Li, J. Chen, N. Li, J. Zheng, J. Zhao, Z. Zhu, Appl. Surf. Sci. 259 Appl. Catal. B: Environ. 163 (2015) 214–222.
(2012) 118–123. [153] Z. Xi, C. Li, L. Zhang, M. Xing, J. Zhang, Int. J. Hydrogen Energy 39 (2014)
[127] J. Chen, X.J. Wu, L. Yin, B. Li, X. Hong, Z. Fan, B. Chen, C. Xue, H. Zhang, 6345–6353.
Angew. Chem. 127 (2015) 1226–1230. [154] M. Jung, J. Scott, Y.H. Ng, Y. Jiang, R. Amal, Int. J. Hydrogen Energy 39 (2014)
[128] J. Ran, J. Zhang, J. Yu, M. Jaroniec, S.Z. Qiao, Chem. Soc. Rev. 43 (2014) 12499–12506.
7787–7812. [155] S. Xu, D.D. Sun, Int. J. Hydrogen Energy 34 (2009) 6096–6104.
[129] O. Rosseler, M.V. Shankar1, M.K. Le Du, L. Schmidlin, N. Keller, V. Keller, J. [156] V. Gombac, L. Sordelli, T. Montini, J.J. Delgado, A. Adamski, G. Adami, M.
Catal. 269 (2010) 179–190. Cargnello, S. Bernal, P. Fornasiero, J. Phys. Chem. A 114 (2010) 3916–3925.
[130] N. Strataki, N. Boukos, F. Paloukis, S.G. Neophytides, P. Lianos, Photochem. [157] A.V. Korzhak, N.I. Ermokhina, A.L. Stroyuk, V.K. Bukhtiyarov, A.E. Raevskaya,
Photobiol. Sci. 8 (2009) 639–643. V.I. Litvin, S.Y. Kuchmiy, V.G. Ilyin, P.A. Manorik, J. Photochem. Photobiol. A
[131] R.M. Navarro, J. Arenales, F. Vaquero, I.D. González, J.L.G. Fierro, Catal. Today 198 (2008) 126–134.
210 (2013) 33–38. [158] J.M. Valero, S. Obregón, G. Colón, ACS Catal. 4 (2014) 3320–3329.
[132] M. Luo, W. Yao, C. Huang, Q. Wu, Q. Xu, J. Mater. Chem. A 3 (2015) [159] J. Yu, Y. Hai, B. Cheng, J. Phys Chem C. 115 (2011) 4953–4958.
13884–13891. [160] W. Wang, S. Liu, L. Nie, B. Cheng, J. Yu, Phys. Chem. Chem. Phys. 15 (2013)
[133] J. Xing, J.F. Chen, Y.H. Li, W.T. Yuan, Y. Zhou, L.R. Zheng, H.F. Wang, P. Hu, Y. 12033–12039.
Wang, H.J. Zhao, Y. Wang, H.G. Yang, Chem. Eur. J. 20 (2014) 2138–2144. [161] W.T. Chen, A. Chan, D.S. Waterhouse, T. Moriga, H. Idriss, G.I.N. Waterhouse,
[134] F.F. Schweinberger, M.J. Berr, M. Döblinger, C. Wolff, K.E. Sanwald, A.S. J. Catal. 326 (2015) 43–53.
Crampton, C.J. Ridge, F. Jäckel, J. Feldmann, M. Tschurl, U. Heiz, J. Am. Chem. [162] N. Naseri, P. Sangpour, S.H. Mousavi, RSC Adv. 4 (2014) 46697–46703.
Soc. 135 (2013) 13262–13265. [163] M. Cheng, M. Zhu, Y. Du, P. Yang, Int. J. Hydrogen Energy 38 (2013)
[135] Y.H. Li, C. Peng, S. Yang, H.F. Wang, H.G. Yang, J. Catal. 330 (2015) 120–128. 8631–8638.
[136] M.C. Hidalgo, M. Maicu, J.A. Navío, G. Colón, J. Phys. Chem. C 113 (2009) [164] R. Su, M.M. Forde, Q. He, Y. Shen, X. Wang, N. Dimitratos, S. Wendt, Y. Huang,
12840–12847. B.B. Iversen, C.J. Kiely, F. Besenbacher, G.J. Hutchings, Dalton Trans. 43
[137] A. Naldoni, M. D’Arienzo, M. Altomare, M. Marelli, R. Scotti, F. Morazzoni, E. (2014) 14976–14982.
Selli, V. Dal Santo, Appl. Catal. B: Environ. 130-131 (2013) 239–248. [165] Z. Jiang, J. Zhu, D. Liu, W. Wei, J. Xie, M. Chen, CrystEngComm. 16 (2014)
[138] V. Jovic, W.T. Chen, D.S. Waterhouse, M.G. Blackford, H. Idriss, G.I.N. 2384–2394.
Waterhouse, J. Catal. 305 (2013) 307–317. [166] E. Qayyum, V.A. Castillo, K. Warrington, M.A. Barakat, J.N. Kuhn, Catal.
[139] C. Wang, D. Astruc, Chem. Soc. Rev. 43 (2014) 7188–7216. Comm. 28 (2012) 128–133.
[140] J.B. Priebe, J. Radnik, A.J.J. Lennox, M.M. Pohl, M. Karnahl, D. Hollmann, K. [167] H. Tian, S.Z. Kang, X. Li, L. Qin, M. Ji, J. Mu, Sol. Energ. Mat. Sol. Cells 134
Grabow, U. Bentrup, H. Junge, M. Beller, A. Brückner, ACS Catal. 5 (2015) (2015) 309–317.
2137–2148.
[141] C.G. Silva, R. Juarez, T. Marino, R. Molinari, H. Garcia, J. Am. Chem. Soc. 133
(2011) 595–602.

You might also like