You are on page 1of 51

Photochemical Water Oxidation: A Review

Submitted by
Kiran Shahzadi

In the partial Fulfillment for the Degree of BS Chemistry

DEPARTMENT OF CHEMISTRY

FACULTY OF SCIENCES

UNIVERSITY OF CENTRAL PUNJAB


(2022)

i
University of Central Punjab
FACULTY OF SCIENCES

Supervisory Committee

The undersigned hereby certify that they have read and recommended to the
Faculty of Sciences for acceptance a thesis entitled “Photochemical Water
Oxidation” by the Kiran Shahzadi Registration# R1F18BSCH0068 is in partial
fulfillment of the requirements for the degree of Bachelor of Science.

Supervisor: _______________
Supervisor Name
Designation
Campus Name
University of Central Punjab

ii
University of Central Punjab
FACULTY OF SCIENCES

Examination Committee

The thesis viva of Kiran Shahzadi (R1F18BSCH0068) was held on date at


Faculty of Sciences, Quaid Campus University of Central Punjab. The supervisory
and Thesis evaluation committee gave satisfactory remarks on the thesis and viva
and were approved for the award of the degree of BS Chemistry.

Thesis Evaluation Committee

1. Name ______________
2. Name ______________
3. Name ______________

__________________

Principal
Campus name
University of Central Punjab

iii
Dedication

To my Parents for their countless love and affection

iv
Acknowledgements
First of all, I am very thankful to ALLAH ALMIGHTY for his limitless blessings. He
bestowed on me that made me able to do this work. It is a great blessing that
MUHAMMAD (S.A.W.W) is my Prophet and all praises are for Him who made
education compulsory for every man and woman.

I would like to express my gratitude to Associate Dean, Faculty of Sciences, Dr.


Naqib Ullah for considering me worthy of writing a thesis and head of department of
Chemistry, Dr. Waqas Shah, for guiding me during my research work. I would like to
pay special thanks to my research supervisor, Dr. Waqas Shah for guiding and assisting
mein whole duration of research work.

I would also like to thank my family and friends for supporting me to reach this
point in my life.
Contents
1. Photochemical Reaction.................................................................................................................. 9
2 Photochemical catalysis ................................................................................................................ 10
3 Water Splitting .............................................................................................................................. 11
3.1 Photochemical water splitting By Oxidation ........................................................................ 12
3.1.1 Mechanistic Aspect for Photocatalytic Water Splitting and a New Strategy to Improve
Activity. 14
3.1.2 Energy requirements ..................................................................................................... 17
3.1.3 Photocatalysts requirements .......................................................................................... 18
3.1.4 Photosensitizers............................................................................................................. 20
3.1.5 Role Of pH in Water Splitting ...................................................................................... 23
3.2 Photochemical Water Splitting by Reduction ...................................................................... 25
3.2.1 Photo-reduction of water to hydrogen using semiconductors and dyes ........................ 26
3.2.2 Role of pH in Photocatalytic Water Reduction: ............................................................ 36
3.2.3 Energy Requirements For Water Reduction ................................................................. 42
3.2.4 Bifunctional water splitting catalyst.............................................................................. 46
FIGURE CAPTIONS

Figure 1: Photochemical Reaction .......................................................................................................... 9


Figure 2: Photochemical Water Splitting .............................................................................................. 13
Figure 3: Evolution Photocataylsts ...................................................................................................... 16
Figure 4: Figure 5: Potential energy diagram for photocatalytic water splitting using a single
semiconductor system (Kudo, 2003) .................................................................................................... 17
Figure 6: Band-gap energy and relative band position of different semiconductors with respect to the
potentials (NHE) for water oxidation/reduction processes. .................................................................. 18
Figure 7: Dye Sensitized Solar Cells .................................................................................................... 21
Figure 8: Sacrificial Agent .................................................................................................................... 21
Figure 9: Scheme of different interactions between different reactive centers in excited state (⁎) or in
ground state. .......................................................................................................................................... 23
Figure 10 The crystal-field-splitting induced electronic configuration of 2H-MoS2 and 1T-MoS2 and
proposed mechanism for catalytic activity of 1T-MoS2. Orange, S; blue-grey, Mo. (b) Time course of
H2 evolved by freshly prepared 1T-MoS2. ............................................................................................ 28
Figure 11 Time course of hydrogen evolution by 1T-MoSe2 and using Eosin Y as a sensitizer. ........ 28
Figure 12 (a) Schematic of process and energy level diagram of hydrogen generation by CdSe–CdS–
Pt heterostructrures. (b) Photocatalytic hydrogen evolution as a function of length of the nanorod and
diameter of the seed. ............................................................................................................................. 30
Figure 13 Schematic illustration of process and energy level diagram of photocatalytic hydrogen
production in the presence of benzyl alcohol........................................................................................ 32
Figure 14 (a) Photocatalytic H2 evolution as a function of Zn substitution in ZnO/Pt/Cd1−xZnxS and
(b) comparison of effect of sacrificial agent on H2 evolution activity. ................................................ 32
Figure 15 (a) High-resolution transmission electron microscopy image of a-TiO2/Pt/CdS and (b) time
course of H2 evolution of a-TiO2/Pt/CdS as a function of time in the presence of benzyl alcohol–
acetic acid.............................................................................................................................................. 33
Figure 16 (a) Comparison of UV–visible absorption spectrum of ZnO1−x(N,F)x with that of undoped
ZnO (inset shows the change in color from white to orange). (b) Comparison of effect of co-doping of
N and F in ZnO on hydrogen evolution of ZnO/Pt/CdS (25%) type hybrid nanostructures using Na2S–
Na2SO3 or benzyl alcohol–acetic acid mixture as sacrificial agent under visible light irradiation. .... 34
Figure 17 Photocatalytic H2 evolution of (a) TiO2/Pt/CdS and (b) ZnO/Pt/CdS heterostructures with
oxides having different surface areas. ................................................................................................... 35
Figure 18 Comparison of photocatalytic hydrogen evolution with (a) different photo catalysts and (b)
ZnO/X/Cd0.8Zn0.2S (X = Pt, NiO and Au). ............................................................................................ 36
Figure 19 Plot of electron-transfer quenching rate (k, = KSv/7) vs. electron-transfer driving force (Eo
= E*Ruw2+ - E(Ru3+-R,,z+ + Ed(Q-d ................................................................................................. 38
Figure 20 pH dependence of the rate of reaction between R~(bpy),~+ and EDTA. ............................ 41
Figure 21 Pathways of light induced hydrogen evolution from water .................................................. 42
Figure 22 Correlation between the observed yields for Hz in the system 4. M Ru(bipy):+, M MV2+, 3.
10-2MEDTAatpH and the radius of the catalysts employed. The lower points at 110 A contain 0.2 mg
of Pt and 0.3 mg of Pt per 25 ml of solution, respectively .................................................................... 43
Figure 23 Procedure for the preparation of ultrafine Pt-sols................................................................. 44
Figure 24 Mechanism of intervention of the ultrafine Pt particles in the H2-formation, electron pool
effect ..................................................................................................................................................... 46
Figure 25 Scheme for cyclic water/decomposition in a combined catalytic system ............................. 47
Figure 26 Sustained light induced water splitting in a system containing bifunctional redox catalyst,
(Ti02/RuOz/Pt) ..................................................................................................................................... 48
Figure 27 Cell device for cyclic water decomposition, oxygen and hydrogen evolution occurs in two
separate half cells .................................................................................................................................. 49
PHOTOCHEMICAL WATER SPLITTING BY
OXIDATION AND REDUCTION
1. Photochemical Reaction
A Photochemical reaction is a reaction which takes place when a chemical process absorbs
light energy as its energy source. The example is the process of photosynthesis. ... The photons
from the light energy split water and enable ATP production that drives the entire process.

Figure 1: Photochemical Reaction

Historical developments in photochemistry took place in the early 1800s. In 1817,


German physicist Theodor von Grotthus developed a theoretical understanding of the
photochemical process. Later, in 1841, American chemist John William Draper studied the
photochemical reaction between hydrogen (H2) and chlorine (Cl2) gases.

A photochemical reaction is based on the principles of photochemistry. When light shines


on a molecule, it goes to an excited state, a process known as photo excitation. There are two
laws of photochemical reaction:

 Grothuss-Draper Law: This law states that a molecule must absorb light in order for a
chemical reaction to take place.
 Stark-Einstein Law: This law states that for each photon of light absorbed by a molecule,
only one molecule is activated for a subsequent reaction.
During a photosynthesis process, the pigment chlorophyll in plants takes in the energy (hν)
from the sun and water (H2O) to convert carbon dioxide (CO2) into glucose (C6H12O6) and
oxygen (O2). Photosynthesis can also be carried in the presence of artificial light.

6CO2 + 6H2O + hν → C6H12O6 + 6O2

Artificial photosynthesis to carry out both the oxidation and the reduction of water has
emerged to be an exciting area of research. It has been possible to photochemically generate
oxygen by using a scheme like the Z-scheme, by using suitable catalysts in place of water-
oxidation catalyst in the Z-scheme in natural photosynthesis. The best oxidation catalysts are
found to be Co and Mn oxides with the e1g configuration.

1 Photochemical catalysis
Photocatalysis is the acceleration of a photoreaction in the presence of a catalyst. In
catalyzed photolysis, light is absorbed by an adsorbed substrate. In photogenerated catalysis,
the photocatalytic activity (PCA) depends on the ability of the catalyst to create electron–hole
pairs, which generate free radicals (e.g., hydroxyl radicals: •OH) able to undergo secondary
reactions. Its practical application was made possible by the discovery of water electrolysis by
means of titanium dioxide (TiO2).

Catalysis in the ground state results from the action of a catalyst, “a substance which
appears in the velocity expression to a power greater than its coefficient in the
stoichiometric equation”. One observes catalysis by a change in rate but explains catalysis by
identifying a new reaction pathway made available by the catalyst. In the simple case, reaction
through the catalytic pathway, which involves a unique set of intermediates and rate constants,
proceeds at a greater rate than does the uncatalyzed reaction and is evidenced in the rate
expression by a term containing the catalyst concentration.

Efficiency for a photochemical reaction is the quantity analogous to rate for a ground-
state reaction. Efficiency is measured by the quantum yield, which is the number of moles of a
particular species formed or reacted divided by the number of Einstein’s of photons absorbed.
In practice, the two measurements comprising the quantum yield are usually rate
measurements, moles per unit of time and Einstein's per unit of time. Rationalization of the
quantum yield, however, in terms of photophysical and photochemical processes, requires a
quantum yield expression, which is analogous to the rate expression for a ground-state reaction.

The quantum yield expression relates efficiency to concentration terms and elementary
rate constants for the system. We propose that a catalyst for a photochemical reaction be
defined as a substance that appears in the quantum yield expression for reaction from a
particular excited state to a power greater than its coefficient in the stoichiometric equation.1°
As will be shown below, it is often convenient to invert the quantum yield expression in order
to isolate the catalyst concentration variable. In the inverted form a catalyst will appear with a
negative exponent, the absolute value of which must exceed its coefficient in the
stoichiometry.(Wubbels, 1983)

2 Water Splitting
Water splitting is the chemical reaction in which water is broken down into oxygen and
hydrogen.

2H2O → 2H2 + O2

Water splitting is a process that enables the production of hydrogen by direct water
decomposition in its elements. The energy required to cleave H—O—H bonds can be supplied
by different power sources: electrical (current), thermal (heat), or light (electromagnetic
radiation). The difference in water splitting processes is made whenever one or another type of
energy source is applied to conduct the reaction, referred to as electrolysis, thermolysis, or
photolysis. Electrolytic water splitting is driven by passing the electrical current through the
water, where conversion of the electrical energy to chemical energy takes place at the electrode-
solution interface through charge transfer reactions in a unit called an electrolyser. Water reacts
at the anode to form oxygen and protons, whereas a hydrogen evolution reaction takes place at
the cathode. Only 3.9% of the world's hydrogen demand is satisfied by electrolysis. Although,
compared with the conventional SMR, the electrolytic water splitting is described as a less
environmentally harmful process with a “zero” CO2 emission (because O2 is the only by-
product), electrolysers powered by the electricity, which is produced by the combustion of coal
or natural gas, resulting in the release of CO2 as a byproduct. Therefore, today's research is
increasingly oriented on utilizing renewable harvesting technologies (wind turbines or
photovoltaics) to drive the electrochemical/catalytic water-splitting reaction. Photochemical
/photocatalytic water splitting is a promising option for hydrogen production which is oriented
on the reduction of CO2-emission and applications of renewable resources such as water and
sunlight. The most important criteria for solar-driven water splitting reactions are the electronic
band gap alignment of the photosensitive material with the redox potential of water. In general,
the presence of transition metal cations with a d0 electronic configuration (Ta5 +, Ti4 +, Zr4 +,
Nb5 +, Ta5 +, W6 +, and Mo6 +), or metal cations with a d10 electronic configuration (In3 +, Sn4 +,
Ga3 +, and Ge4 +) is important for the efficient photocatalytic materials, the empty d or sp
orbitals of which form the bottom of the respective conduction bands. Over the last decades,
considerable progress in this field has been made by an increasing number of research groups
where the topic was thoroughly reviewed.

Thermal or thermochemical water splitting is another alternative technology to produce


hydrogen from water with potentially low or no greenhouse gas emissions. This technology
has been extensively studied by many research groups, and more than 300 water-splitting
cycles with different operating parameters, engineering challenges, and hydrogen production
opportunities are described in the literature. Thermochemical water splitting processes require
hot temperatures (500–2000°C) to drive a series of chemical reactions, which lead to hydrogen
evolution. The chemicals used in the process are reused within each cycle, creating a closed
loop that consumes only water and produces hydrogen and oxygen. This technology is an
appealing pathway to produce hydrogen-utilizing waste heat from existing nuclear power
stations or concentrated (using a field of mirror “heliostats”) solar power. However, a
realization of renewable-energy solutions for the water splitting reactions on a larger scale
remains a challenge, which is dictated by the overall economy of the process and heavily
depends on further developments of cost-effectiveness and environmentally benign
technologies.(Maeda and Domen, 2010)

2.1 Photochemical water splitting By Oxidation

Photocatalytic overall water splitting to form hydrogen and oxygen has attracted
considerable attention as a potential means of renewable energy production with no reliance
on fossil fuels and no carbon dioxide emission.1-5 as illustrated in Figure 1, current successful
photocatalytic systems for overall water splitting can be divided into two primary approaches.
One approach is to split water into H2 and O2 using a single visible-light-responsive
photocatalyst with a sufficient potential to achieve overall water splitting. In this system, the
photocatalyst should have a suitable thermodynamic potential for water splitting, a sufficiently
narrow band gap to harvest visible photons, and stability against photo corrosion. Because of
these stringent requirements, the number of reliable, reproducible photocatalysts suitable for
one-step water splitting is limited.6, 7 the other approach is to apply a two-step excitation
mechanism using two different photocatalysts.8 this was inspired by natural photosynthesis in
green plants and is called the Z-scheme. The advantages of a Z-scheme water splitting system
are that a wider range of visible light is available because a change in Gibbs free energy
required to drive each photocatalyst can be reduced as compared to the one-step water splitting
system and that the separation of evolved H2 and O2 is possible. It is also possible to use a
semiconductor that has either a water reduction or oxidation potential for one side of the
system. For example, some metal oxides (e.g., WO3 and BiVO4) function as a good O2
evolution photocatalyst in a two-step water splitting system using a proper redox mediator,
although they are unable to reduce water. Successful overall water splitting via two-step photo
excitation by visible light using several combinations of photocatalysts and electron relays has
been reported. However, challenges remain in the promotion of electron transfer between two
semiconductors and in the suppression of backward reactions involving shuttle redox
mediators. As we can expect from the reaction scheme shown in Figure 2.

Figure 2: Photochemical Water Splitting

Photocatalytic activity for overall water splitting is strongly dependent on the


physicochemical properties of a photocatalyst, the nature of the active sites (so-called
cocatalyst), and the reaction conditions.2,4 In the past decade (in particular, from 2000 to
2005), a number of materials have been reported as visible-light-driven photocatalysts capable
of producing both hydrogen and oxygen under visible light.1-4,17 Some have successfully
achieved overall water splitting without any sacrificial reagents. In the past 5 years,
considerable progress has been made on cocatalyst development and the elucidation of reaction
mechanisms. This Perspective highlights some important aspects of recent water splitting
research.(Youngblood et al., 2009)

2.1.1 Mechanistic Aspect for Photocatalytic Water Splitting and a


New Strategy to Improve Activity.

During the development of photocatalytic systems for overall water splitting under visible
light, several photocatalyst materials and preparation methods have been reported. Many
studies have focused on the development of materials that are suitable for visible-light-driven
overall water splitting by addressing light absorption properties, band edge position,
crystallographic quality, particle morphology, and phase purity. However, it is difficult to
understand what factor(s) dominates the net photocatalytic activity based on the above physical
properties because the photocatalytic reactions proceed through a complicated sequence of
competing multistep processes. This demonstrates the importance of understanding the kinetics
and dynamics of a photocatalytic reaction to establish rational strategies for the immediate
development of photocatalytic systems and for future practical applications. Aspects of the
photocatalytic water splitting mechanism on GaN:ZnO powder modified with a Rh-Cr mixed
oxide cocatalyst were examined with respect to the effects of cocatalyst loading, light intensity,
hydrogen/deuterium isotopes, and reaction temperature on photocatalytic activity.27 The water
splitting rate with the optimally modified photocatalyst was proportional to light intensity
under solar-equivalent or weaker irradiation, indicating that accumulation of photoexcited
electrons and holes was negligible. Excess loading of the cocatalyst did not improve the water
splitting rate. The H-D isotope effect on overall water splitting was significantly lower than
previously reported values for photocatalytic and electrochemical H2 evolution reactions. The
apparent activation energy for overall water splitting was as low as 8 kJ mol-1 and was
unchanged by the addition of electron donors or acceptors. These results reflect a shortage of
photoexcited carriers available for surface redox reactions under steady light irradiation. In
summary, the experimental results indicate that the balance between the rates of redox reactions
on the photocatalyst surface and carrier generation/ recombination in the photocatalyst bulk
determines the steady state charge concentration in the photocatalyst, that is, developing both
a photocatalyst and a cocatalyst is important. The proposed kinetic model of photocatalytic
water splitting also suggests that the reaction probability of photoexcited holes for O2 evolution
versus recombination with intrinsic electrons in the photocatalyst determines the water splitting
activity of GaN:ZnO. It would be natural to expect that loading both H2 and O2 evolution
cocatalysts onto the same photocatalyst would improve water splitting activity, compared to
that for photocatalysts modified with either a H2 or O2 evolution cocatalyst. It is easy to
imagine that the two different cocatalysts would separately facilitate H2 and O2 evolution,
thereby promoting overall water splitting in harmony. Unfortunately, no successful, reliable
example of this has been reported since the initial reports on photocatalytic water splitting in
the 1980s, and a demonstration of the concept has remained a challenge. Very recently, we
demonstrated a proof of concept using GaN:ZnO loaded with Rh/Cr2O3 (core/shell) and
Mn3O4 nanoparticles as H2 and O2 evolution promoters, respectively, under visible light
irradiation (λ > 420 nm). 28 The goal was to generate both H2 and O2 evolution sites separately
on the same photocatalyst surface. The preparation method developed by our group is a
stepwise deposition involving the adsorption of MnO nanoparticles (9.2 (0.4 nm) followed by
calcination to give crystallized Mn3O4 nanoparticles and a subsequent photo deposition of
Rh/Cr2O3 (core/shell) nanoparticles, as shown schematically in Figure 4A. This procedure
allowed the separate construction of H2 and O2 evolution sites, as revealed by TEM
observation and energy dispersive X-ray spectroscopy analysis. As mentioned earlier,
core/shell structured Rh/Cr2O3 nanoparticles provide active sites for H2 evolution. On the
other hand, photoelectrochemical analysis revealed that Mn3O4 nanoparticles on GaN:ZnO
promote photooxidation of water. Finally, overall water splitting was attempted using the as-
prepared samples under visible light. As expected, the activity of GaN:ZnO modified with both
Rh/Cr2O3 and Mn3O4 provided a higher activity than modification with either Rh/Cr2O3 or
Mn3O4, as shown in Figure 4B. This demonstrates the validity of the above idea and suggests
a new strategy for improving photocatalytic activity for overall water splitting. The proposed
reaction scheme of overall water splitting on GaN:ZnO modified with Rh/Cr2O3 and Mn3O4.
Prospects for Photocatalytic Overall Water Splitting. As described above, in the search for
visible-light responsive photocatalysts, significant effort has been devoted to the development
of active sites on photocatalysts and elucidating reaction mechanisms, leading to significant
progress in the field of heterogeneous photo catalysis for water splitting, especially in the last
5 years.18-29 We have developed several promising systems, including Rh2-yCryO3- loaded
GaN:ZnO (one-step water splitting system) and a two-step system consisting of Pt/ZrO2/TaON
and Pt/WO3 with an IO3 -/I- shuttle redox mediator, with respective apparent quantum yields
of about 5.1% at 410 nm and 6.3% at 420.5 nm. However, we continue to pursue more active
photocatalytic systems capable of harvesting more visible photons. As shown in Figure, solar
energy conversion efficiency increases when one can achieve overall water splitting under
longer wavelength irradiation. This is because the number of available photons in solar
spectrum increased with increasing wavelength.(Maeda and Domen, 2010)

Figure 3: Evolution Photocataylsts

To provide one-third of the projected energy needs of human society in 2050 from solar
energy, our preliminary estimation suggests that approximately 10 000 “solar plants” (5 km 5
km in area per plant) with a solar energy conversion efficiency of 10% would be needed. The
total required area, 250 000 km2, corresponds to 1% of the earth's desert area; 570 tons of H2
gas would be produced per day, assuming an integrated solar energy of AM1.5G irradiation for
a day with correction for sunlight angle. This H2 would be available for use as a “recyclable”
reactant in fuel cells and as a raw material to produce important chemicals such as methanol
and so on. Of course, a technology to separate simultaneously produced H2 and O2 would be
required. This scheme is illustrated in Figure. In relation to such a system processing, a unique
reactor toward photoelectrochemical hydrogen production was proposed recently. James et al.
reported a technoeconomic evaluation of conceptual photochemical H2 production systems
using solar energy. According to that report, shallow horizontal pools or beds consisting of a
flexible clear plastic thin-film baggie, which contains a reactant solution and a photocatalyst,
would be a potential candidate as an inexpensive water splitting reactor. Such a system-
processing study, including the construction of the water splitting reactor, H2 and O2 gas
separator, a solar hydrogen chemical plant, and so on, is expected to be more important for
realizing practical application in the future. Requirements for Photochemical Water

Splitting

Splitting of water to its constituents (i.e., hydrogen and oxygen) for production of hydrogen
energy at an industrial scale is one of the "holy grails" of materials science. That can be done
by utilizing the renewable energy resource i.e., sunlight and photocatalytic material. The
sunlight and water are abundant and free of cost available at this planet. But the development
of a stable, efficient and cost-effective photocatalytic material to split water is still a great
challenge. To develop the effective materials for photocatalytic water splitting, various type of
materials with different sizes and structures from nano to giant have been explored that includes
metal oxides, metal chalcogenides, carbides, nitrides, phosphides, and so on.

2.1.2 Energy requirements

Energy requirements for photochemical water reduction to occur, the flat band potential
of the photocatalytic semiconductor must exceed the proton reduction potential (0.0 V against
the normal hydrogen electrode, NHE, at pH = 0,). Furthermore, to facilitate water oxidation,
the VB edge must exceed the oxidation potential of water (+1.23 V against the NHE, at pH =
0). Therefore, according to these values, a theoretical semiconductor band-gap energy of 1.23
eV is required to drive the overall water-splitting reaction according to Equation:

H2O H2+ 1/2O2

Figure 4: Figure 5: Potential energy diagram for photocatalytic water splitting using a single semiconductor
system (Kudo, 2003)
2.1.3 Photocatalysts requirements

Photocatalysts requirements for the photodissociation of water must satisfy several functional
requirements with respect to semiconducting and electrochemical properties:

i. Suitable solar visible-light absorption capacity with a band gap of around 2.0–2.2 eV
and band edge potentials suitable for overall water splitting,
ii. Capacity for separating photoexcited electrons from reactive holes,
iii. Minimization of energy losses related to charge transport and recombination of
photoexcited charges,
iv. Chemical stability opposed to corrosion and photo corrosion in aqueous environments,
v. Kinetically suitable electron transfer properties from photocatalyst surface to water, and
vi. Being easy to synthesize and cheap to produce.

The basic parameter deciding the light-harvesting ability of the photocatalyst is its electronic
structure that determines the band-gap energy. Figure 10 illustrates the band positions of
various semiconductors regarding the potentials (NHE) for water-oxidation/reduction
processes (Xu and Schoonen, 2000). From the perspective of band positions, among the
semiconductors represented in Figure 10, those that fulfil the thermodynamic requirements for
overall water splitting are KTaO3, SrTiO3, TiO2, ZnS, CdS, and SiC.

Figure 6: Band-gap energy and relative band position of different semiconductors with respect to the potentials
(NHE) for water oxidation/reduction processes.
However, it is important to pressure that the potential of the band structure of the
semiconductor is just the thermodynamic requirement. There is an activation wall in the charge
transfer process between photocatalyst and water molecules derived from the energy losses
associated with solar energy adaptation on photocatalysts: thermodynamic losses, transport of
electron/holes, electron/holes recombination, and kinetic losses. The existence of these energy
losses rises the optimum band gap for high-performance photocatalysts from the theoretical
value of 1.23 to 2.0–2.2 eV (Bolton, 1996). Another essential requirement for the photocatalyst
is its opposition to reactions at the solid/liquid interface that may result in a degradation of its
properties. These reactions include electrochemical corrosion, photo corrosion, and dissolution
(Morrison, 1980). A great group of photocatalysts with suitable semiconducting properties for
solar energy conversion (CdS, GaP, etc.) are not stable in the water-oxidation reaction because
the anions of these materials are more susceptible to oxidation than water, causing their
degradation by oxidation of the material (Ellis et al., 1977; Williams, 1960). Water-splitting
reactions at the photocatalyst interface can occur if charge carriers generated by absorbed light
can reach the solid–liquid interface during their lifetime and are capable of finding suitable
reaction partners – protons for electrons and water molecules for holes. For that reason, the
generation and separation of photoexcited carriers with a low recombination rate is also an
essential condition to be fulfilled by the photocatalysts. The transport of photoexcited carriers
strongly depends on both the microstructural and surface properties of photocatalysts. In
general, the high crystalline level of the photocatalyst has a positive effect on photoactivity, as
the density of defects caused by grain boundaries, which act as recombination centers of
electrons and holes, decreases when particle crystallinity increases (Ikeda et al., 1997;
Kominami et al., 1995; Reber and Meier, 1984). Surface properties such as surface area and
active reaction sites are also important. The surface area, determined by the size of the
photocatalyst particles, also influences the efficiency of charge carrier transport.(Barendrecht,
1982) To have an efficient charge transport, the diffusion length of charge carriers must be long
compared to the size of the particles. Therefore, the possibility of the charge carrier reaching
the surface increases as the size of the photocatalysts decreases (Ashokkumar, 1998).
Nevertheless, the improvement in the efficiency associated with the high crystalline level of
the photocatalyst prevails over the improvement associated with small-sized particles. (Tee et
al., 2017)
2.1.4 Photosensitizers

Photosensitizers are molecules which absorb light (hν) and transfer the energy from
the incident light into another neighboring molecule. This light is frequently within the visible
spectrum or infrared spectrum, as any higher energy electromagnetic radiation may result in
the photoelectric effect. Photosensitizers experience varying levels of efficiency for
intersystem crossing at different wavelengths of light based on the internal electronic structure
of the molecule.

For a molecule to be considered a photosensitizer:

 The photosensitizer must impart a physicochemical change upon a substrate after


absorbing incident light.
 Upon imparting a chemical change, the photosensitizer returns to its original chemical
form.

Photosensitizers have existed within natural systems for as long as chlorophyll and other
light sensitive molecules have been a part of plant life, but studies of photosensitizers began as
early as the 1900s, where scientists observed photosensitization in biological substrates and in
the treatment of cancer. Mechanistic studies related to photosensitizers began with scientists
analyzing the results of chemical reactions where photosensitizers photo-oxidized molecular
oxygen into peroxide species. The results were understood by calculating quantum efficiencies
and fluorescent yields at varying wavelengths of light and comparing these results with the
yield of reactive oxygen species. However, it was not until the 1960s that the electron donating
mechanism was confirmed through various spectroscopic methods including reaction-
intermediate studies and luminescence studies.

Photosensitizers can be retained into 3 generalized domains based on their molecular


structure. These three fields are organometallic photosensitizers, organic photosensitizers, and
nanomaterial photosensitizers. Via the absorption of light, photosensitizers can consume triplet
state transfer to reduce small molecules, such as water, to generate Hydrogen gas. As of right
now, photosensitizers have generated hydrogen gas by splitting water molecules at a small,
laboratory scale.
Figure 7: Dye Sensitized Solar Cells

Dye sensitized solar cells are photosensitizers which transfer energy to semiconductors
to generate energy from solar light.(Meza-Chincha et al., 2021)

2.1.4.1 Sacrificial Electron Donors


Sacrificial agents are the electron donors or slum hunters that reduce the
recombination tendency of electrons and holes and quicken the rate of hydrogen generation.

Figure 8: Sacrificial Agent

Hydrogen generation over carbon-, nitrogen and sulfur-doped TiO2 semiconductor


photocatalysts (represented as C–TiO2, N–TiO2 and S–TiO2, respectively) under visible light
irradiation had been achieved using various sacrificial electron donors, namely
triethanolamine, diethanolamine, monoethanolamine, triethylamine, MeOH, EtOH, EDTA, L-
ascorbic acid and phenol. The highest initial rate of H2 production was originate to be in the
range 1,000–2,200 µmol/g/h at ambient conditions when triethanolamine was used as
sacrificial electron donor. The efficiency of hydrogen production over these photocatalysts
depends strongly on the nature of the sacrificial electron donor and falls in the following order:
C–TiO2 > S– TiO2 > N–TiO2. The results of the present studies suggest that the rate of H2
production is not simply governed by the reduction potential of the sacrificial electron donor
but also by the kinetic hurdle of the electron transfer process.(Schneider and Bahnemann, 2013)

2.1.4.2 Sacrificial Electron Acceptor


Similarly to any other redox cycle, for allowing the transits of the electron towards the
chemisorbed CO2, the semiconductor must “purchase” electrons from sacrificial agent,
guaranteeing the electronic neutrality of the material. In particular, the oxidation of water to
O2 is the main way for providing to the wanted electrons, Therefore, the methanol
synthesis involves the reduction of electron “acceptor” reactants (i.e., CO2, H2O, etc.) to form
methanol, H2, CO, and other organic compounds, and the oxidation of donor species (i.e., H2O,
and OH−), through the combined action of holes and electrons which move on the surface of
the photocatalyst (Ola and Maroto-Valer, 2015c).

As well known, the initial excitation, which occurs in any semiconductor during the
photocatalytic process if properly photo-stimulated, drives to electron transfer and/or energy
transfer phenomena as a natural consequence of the converse process of de-excitation. In
respect of this, electron transfer represents the driving power of any chemical reaction which
occurs on the surface of the heterogeneous photocatalyst. Basically, the photoactivity involves
specific “reactive centers”, which are reactive sites placed on the surface of the heterogeneous
photocatalyst, where the electrons jump from the occupied orbital of the donor species to the
empty orbital of the acceptor reactants.

As shown in the scheme of Figure, the electron transfer process occurs through the
overlap of the orbitals of the donor and acceptor species, also partially filled. This process is
realized through the relocation of an electron from the donor to the acceptor, leading to the
formation of cation (D+) and anion (A−) species. Differently, the energy transfer occurs by the
phenomena of electron exchange or dipole-dipole resonant coupling. If equally favored by the
thermodynamic, the process of electron transfer results larger than that of the electron
exchange, because the latter needs the concomitant overlap of two orbital pairs. While the
dipole-dipole coupling, occurring via Coulombic resonance interactions, is easily achieved
because this process doesn't involve a real orbital overlapping and also acts in a wider range of
10–100 Å. (Zhao et al., 2020)
Figure 9: Scheme of different interactions between different reactive centers in excited state (⁎) or in ground state.

2.1.5 Role Of pH in Water Splitting

Solar water splitting is an attractive approach to overcome the intermittency of sunlight


and store it as hydrogen. For a large-scale, stable and safe operation, the use of near-neutral pH
solutions is preferred. However, although an efficient (photo) electrocatalyst exists for near-
neutral pH operation, the mass-transport limitation of proton and hydroxide ions is a major
challenge. Understanding this limitation and how it leads to the formation of pH gradient and
the resultant efficiency loss is highly important to devise an efficient design of energy
conversion devices, even beyond water splitting (e.g., CO2 reduction). Herein, pH-sensitive
fluorescence sensor foils are introduced in an electrochemical cell to monitor the local pH in
situ during a water splitting reaction in neutral pH electrolytes. A local pH shift and the
appearance of natural convection are clearly visualized. Combined with a multiphysics model,
our results show that the natural convection is driven by the change of electrolyte density and
plays a significant role to suppress local pH shift. This factor, which is not previously
highlighted, should be considered when determining the different design parameters (e.g., flow
rate, the choice of anions and cations, product separation strategy) of a (photo) electrochemical
device.(Najafi et al., 2019)

2.1.5.1 Hydrogen Evolution Reaction:


HER, as a vital half reaction of electrochemical water splitting, proceeds a 2e reaction
on the surface of cathode electrode. In general, two steps are included for H2 evolution. Firstly,
H+ is discharged to couple with one e on catalytic sites to form an intermediate (*H), which is
viewed as Volmer step. Then, H2 is produced from two different processes, Heyrovsky or Tafel
steps [129–131]. For Heyrovsky step, one *H reacts with both an H+ and an e_ to generate H2.
For Tafel step, two *H are combined to form H2. Thus, HER in acid and alkaline electrolytes
can be categorized to two reaction mechanisms (Volmer-Heyrovsky and Volmer-Tafel
mechanisms, equations (1)–(4)).

In acid:

H+ + * + e _ → *H; *H + H+ + e_ → * + H2 (1)

H+ + * + e- → *H; 2*H → 2* þ H2 (2)

In alkali:

H2O + * + e_ → *H + OH- ; *H + H2O + e-→ * + H2 + OH- (3)

H2O + * + e- → *H + OH- ; 2*H → 2* + H2 (4)

For HER, hydrogen adsorption Gibbs free energy (ΔGH*) is an indicator for evaluating
reaction rate. If ΔGH* on the surface of electrocatalyst is negative, the Volmer step will occur
easily, and the Heyrovsky step or Tafel step will be regarded as the rate-determining step of
HER. However, the whole HER will be difficult to occur when ΔGH* is positive. In addition,
the dissociation energy of H2 on the surface of electrode catalysts is also a vital factor to
evaluate HER rate. A value of ~0 is optimal for ΔGH*. For example, the value of ΔGH* on
Pt(1 1 1) is close to 0 in acidic conditions, which is highly active for HER.

2.1.5.2 Oxygen Evolution Reaction


OER is an oxidation reaction in electrochemical water splitting, which occurs at anode through
4e pathways (equations (5)–(12)).

H2O + * → *OH + H+ + e- (5)

*OH → *O + H+ + e- (6)

*O + H2O → *OOH + H+ + e- (7)

*OOH → * + O2 + H+ + e- (8)

In alkali:

OH + * → *OH + e- (9)

*OH + OH- *O + H2O + e (10) *O + OH → *OOH + e (11) *OOH + *OH → * O2 +


H2O + e (12) As compared to HER, OER proceeds more complicated steps, and various
intermediates (including *OH, *O, and *OOH species) are produced in acidic and alkaline
media [6]. Due to the complicated multi-step reactions of OER, high OER over potential should
be conquered. Therefore, developing high-performance electrocatalysts to lower OER
overpotential is crucial. For OER, the reaction rate can be evaluated by the changes of Gibbs
free energy between adjacent intermediates (ΔG), which can be tuned to the optimal over
potentials [9]. Additionally, the volcano plots can provide a guide for the design of
electrocatalysts, by which the structure of electrocatalysts is able to be adjusted to change the
binding energy of intermediates on catalytic sites, further realizing the rapid OER kinetics. For
example, Ru element is located at the top of volcano plot for OER, indicating that superior
OER performance can be realized on Ru-base. (Najafi et al., 2019)

2.2 Photochemical Water Splitting by Reduction

Increasing global claim for energy and disquiets over anthropogenic climate change
drive the search for sustainable, another energy sources. Solar energy is a leading candidate, as
it achieves the scale needed to meet the demand, but it suffers from serious problems in energy
storage and transport. One striking solution to this problem is solar-to-fuel conversion, where
energy is stored by changing energy-poor molecules into energy-rich ones. In particular,
hydrogen is a pleasing carbon-neutral fuel, as both its feedstock and combustion product are
water. Platinum can catalyze the reduction of water to hydrogen at fast rates near
thermodynamic equilibrium, but its low profusion and high cost forbid its widespread use. As
such, it is striking to develop cheap, earth-abundant hydrogen-evolution catalysts that can,
when interfaced with electrodes or photo electrodes, convert solar energy into chemical energy
in a viable manner. In this context, biology and materials provide fusing concepts to drive the
development of new solar-fuel chemistry. Within the configuration of hydrogen production,
hydrogenase enzymes can convert protons and electrons to hydrogen consuming only iron
and/or nickel in their active sites with high speed and efficacy on a per molecule basis.
However, the activity of enzymes on a per volume basis is limited by size limits, as most of the
protein structure is dedicated to its biological regulation for cellular function while serious
catalytic reactions occur at relatively small active sites within these larger frames. A similar
situation is observed in heterogeneous materials, where catalytically active sites are often
constrained to minor edges or faces and the bulk of the material is inert and provides but a
framework for stability purposes. With these principles in mind, molecular catalysts can
maintain the small, functional units of their biological or materials counterparts but can achieve
higher volumetric activities by leaving the excessive regulatory or stabilizing frameworks.
Indeed, molecular proton reduction catalysis is an exciting area of research and has been the
subject of many excellent and comprehensive reviews.

2.2.1 Photo-reduction of water to hydrogen using semiconductors


and dyes

Photocatalysts such as TiO2, SrTiO3 and ZnO possess desirable VBs and CBs, but they
are only energetic under UV light due to the large band gaps. The best method to make them
visible light-responsive is by doping other elements, forming solid solutions or sensitization
with other materials (dyes, small band gap materials). For example, forming a solid solution of
ZnS with AgInS2 as in (ZnS)0.4(AgInS2)0.6 results in the optimum band gap with appropriate
band positions. These nanocrystals capped by S2− ions are produced by ligand exchange
reaction in toluene–formamide solvent intermediate. They show photocatalytic H2 evolution
activity of 5.0 mmol g−1 h−1 without the addition of any noble metals which is considerably
higher compared with its bulk counterpart (1.1 mmol h−1 g−1).i Doping of anions (donors) in
oxides and sulfides is a more capable approach than the doping of cations (acceptors) to bring
about substantial changes in the electronic structure. Substitution of O with N in these materials
alters the electronic structure and falls the absorption edge. However, it leads to charge
imbalance due to interchanging O2− with N3− and defective. Co-substitution of N and F in place
of O is a promising method to alter the electronic structure while retaining the charge balance.ii
Lately, co-substitution of N and F in TiO2 and ZnO has been carried out in our laboratory which
leads to extensive decrease in band gaps of TiO2 and ZnO with a change in color from white
to yellow and orange, correspondingly.iii TiO2−x(N,F)x exhibited a hydrogen evolution rate of
60 μmol h−1 g−1 without the occurrence of any noble-metal catalyst under visible light
irradiation.iv Hence co-substitution of N and F in several wide band gap oxides such as SrTiO3,
NaTaO3, BaTaO3, etc., is a potential method to create and use them as visible light-sensitive
photocatalysts.

2.2.1.1 (a) Dyes

photocatalyst dye activity (mmol g−1 h−1)

2H-MoS2 eosin 0.05 (0.008)


MoS2-NRGO eosin 10.8 (2.9)

1T-MoS2 eosin 30 (6.5)

1T-MoSe2 eosin 60 (19.2)

R-HCa2Nb3O10 eosin 4.3

7% Pt/g-C3N4 eosin 1.6

7 wt% Pt/RGO eosin 4.6

Dyes have been successfully used as visible light-responsive components in


photocatalysis. Combinations of dyes with varied variety of materials, ranging from large band
gap to small band gap and metallic systems have been studied (table).v For example, in dye-
assisted TiO2, the dye absorbs light and inserts the photo-excited electron to the CB of the
TiO2 and these electrons participate in the reduction of H+ to H2. Use of small band gap
materials such as MoS2 with Eosin Y dye has been studied lately.vi Hydrogen evolution
activities of 0.05 and 2–3 mmol h−1 g−1 were obtained with 2H-MoS2 and graphene–MoS2,
respectively. Graphene acts as an electron channel in the latter situation. Use of N-doped
graphene with MoS2 results in the development of p–n junction which results in the induced
electric field at the interface of n-type N-doped graphene and p-type MoS2. As a result,
enhanced hydrogen evolution action of 10.8 mmol h−1 g−1 is obtained (table). 1T-MoS2 exhibits
an activity of 30 mmol h−1 g−1, superior to those achieved with 2H-MoS2. In figure as it is
accessible the mechanism of dye-assisted photocatalytic hydrogen evolution. Upon visible
light irradiation EY forms EY1* and then transforms to triplet EY3*. It gets electrons from
sacrificial electron donor and forms the EY− anion. The electron from the anion is shifted to
MoS2 (1T, metallic form) either openly or through graphene channel. Recently, 1T and 2H
arrangements of MoSe2 have also been used along with Eosin Y for visible light-induced
hydrogen evolutionvii. 1T-MoSe2 exhibits an activity of 60 mmol h−1 g−1, higher to those
obtained with 2H-MoSe2 (0.05 mmol h−1 g−1), 1T-MoS2 (30 mmol h−1 g−1) and 2H-
MoS2 (0.05 mmol h−1 g−1) with a turnover factor of 19.2. In table, these results are compared
with those obtained with alike other systems. viii
Figure 10 The crystal-field-splitting induced electronic configuration of 2H-MoS2 and 1T-MoS2 and proposed mechanism for
catalytic activity of 1T-MoS2. Orange, S; blue-grey, Mo. (b) Time course of H2 evolved by freshly prepared 1T-MoS2.

Figure 11 Time course of hydrogen evolution by 1T-MoSe2 and using Eosin Y as a sensitizer.
2.2.1.2 (b) Semiconductors
Very few of the narrow band gap semiconductors hold desirable CB and VB positions.
Among them CdS, CdSe (quantum dots), TaON, Y2Ta2O5N2, etc., have been well discovered.
Recently, g-C3N4 has increased attention owing to its desirable VB and CB positions in addition
to its visible light absorption and surface catalytic sites Alteration of the electronic structure of
these materials, for example P and Cl co-substituted CdS, has been shown to enhance the
activities approximately three times probably due to the capability for using longer
wavelengths. These materials themselves display very weak H2 evolution activities because of
the recombination of either charge carriers or the reaction intermediates. Often redox reactions
of water splitting arise on the time scales of microseconds or longer, whereas the charge carrier
recombination and relaxation happen within a few nanoseconds. In order to achieve maximum
yields, there is a need to extend the lifetime of charge carriers and suppress the recombination
of reaction intermediates.

Kato et al have demonstrated efficient water splitting to hydrogen and oxygen using
NiO–NaTaO3: La by physical departure of water reduction and oxidation sites. This has
motivated several researchers to take steps to develop such kinds of structures and estimate
their properties. A very important contribution has been made by Alivisatos and co-workers
who have developed multicomponent nanoheterostructures with tunable properties. These
structures contain Pt-tipped CdS nanorod with fixed CdSe seed. Upon photo excitation of CdS,
the photo-generated electrons are inserted into Pt and holes are confined to CdSe seeds. These
authors could tune the properties by change the distance between the reactions sites and extent
of charge separation through band alignment of CdS and CdSe. From this study, it was
concluded that growing nanorod length (i.e. the distance between the reaction sites, Pt and
CdSe) results in increased hydrogen evolution activities. Similarly, a smaller seed (2.3 versus
3.1 nm) results in greater yields. CdSe/CdS/Pt heterostructures with seed diameter of 2.3 nm
and nanorod length of 60 nm displayed a hydrogen evolution of 40 mmol h−1 g−1 with an
apparent quantum yield (AQY) of 20% at 420 nm in the presence 2-propanol (table). And their
study also exposed that hydrogen bound to Pt is stable and the reaction is of first order with
respect to the photon flux.
Figure 12 (a) Schematic of process and energy level diagram of hydrogen generation by CdSe–CdS–Pt heterostructrures. (b)
Photocatalytic hydrogen evolution as a function of length of the nanorod and diameter of the seed.

Table 1

photocatalyst co- activity AQY reference


catalyst (mmol h−1g−1) (%)

CdS–CdSe– Pt 40 20 [36]
Pt
ZnO–Pt– Pt 32 36 [37]
CdS
ZnO–Pt– Pt 37 50 [37]
Cd0.8Zn0.2S
a- Pt 40 48 this study
TiO2/Pt/Cd0.8Zn0.2S
CdS– Pt 56 23b [38]
graphene–Pt
CdS/(Pt– Pt 9 5 [39]
TiO2)
CdS/Pt/TiO2 Pt 15 14b [40]

a
Visible light irradiation.

b
Light sources used for activity and AQY calculations are different; AQY, apparent
quantum yield.
Similar strategies have been used lately, by forming heterostructures of several other
semiconductors. Heterostructures of ZnO/CdS and TiO2/CdS have been well discovered. Band
positions in these structures form a type II alignment which is fit for charge separation.
Therefore, higher hydrogen evolution activities are obtained with ZnO/CdS and TiO 2/CdS
relative to their individual components. Deposition of co-catalysts such as RuO2 and Pt delivers
active sites for surface redox reactions and improves the activity. Wang et al have saw that the
charge carriers in ZnO/CdS exhibit lifetimes up to 220 ns, whereas in ZnO and CdS have only
60 ns decay lifetime.

We have lately developed simple and solution-processed ZnO/Pt/CdS-type hetero-


structures. In these heterostructures, upon visible light irradiation CdS absorbs the light and
makes electron–hole pairs. The electrons are injected to ZnO and ultimately reach Pt and reduce
H+ to H2. The holes found in the VB of CdS are used by the sacrificial agents. The presence of
Pt in these structures shows about sevenfold improvement in H2 evolution activity. The
existence of Pt on ZnO is twofold more highly active compared with Pt present on CdS,
indicating the importance of location of Pt. In this configuration of ZnO/Pt/CdS, vectrorial
transferal of excited electrons makes them suitable for charge transfer and separation. Partial
substitution of Zn in CdS as in ZnO/Pt/Cd1−xZnxS has shown amazing changes in the activities.
ZnO/Pt/Cd0.8Zn0.2S heterostructures have shown a maximum hydrogen evolution of 12.5 and
31.2 mmol h−1 g−1 in the presence of Na2S–Na2SO3 under visible and UV–visible irradiation,
respectively. Although the substitution of Zn in CdS increases the band gap making fewer
photons fit for absorption, we see enhanced H2 evolution activity. This could be due to the
improved charge separation obtained by altering the band alignment by partial substitution of
Zn in CdS. This fact was also sustained by room temperature photoluminescence studies.
Similar effect has also been saw with partial substitution of Se in CdS as in ZnO/Pt/CdS1−xSex.
On changing Na2S–Na2SO3 with benzyl alcohol–acetic acid, ZnO/Pt/CdS and
ZnO/Pt/Cd0.8Zn0.2S showed nearly threefold improvement in activity yielding 31.6 and
36.5 mmol h−1 g−1, correspondingly, under visible light irradiation. In table, we match the
yields obtained by these systems with other similar systems.
Figure 13 Schematic illustration of process and energy level diagram of photocatalytic hydrogen production in the presence
of benzyl alcohol.

Figure 14 (a) Photocatalytic H2 evolution as a function of Zn substitution in ZnO/Pt/Cd1−xZnxS and (b) comparison of effect
of sacrificial agent on H2 evolution activity.

We have carried out a detailed investigation of the visible light-induced hydrogen evolution by
using TiO2/Pt/Cd1−xZnxS (x=0.0, 0.2) heterostructures, wherein poorly crystalline or nearly
amorphous TiO2 (anatase) as well as crystalline anatase (denoted as a-TiO2 and c-TiO2,
respectively) have been used. a-TiO2 was prepared by the hydrolysis of aqueous TiCl4 with
dilute KOH solution and c-TiO2 was prepared by thermal treatment of a-TiO2 at 300°C. With
the increasing proportion of CdS, the absorption in the visible region increases as expected.
We first studied visible light-induced hydrogen evolution in the presence of Na2S–Na2SO3 as
sacrificial agents and found maximum activity at a loading of 20 mol% of CdS relative to
TiO2 in a-TiO2/CdS. The fraction of CdS that effectively forms the interface with the
TiO2 decreases beyond 20 mol%. This reduces the transfer of excited electrons to the TiO2 and
also prevents the reactant molecules from reaching the TiO2 surfaces. We have, therefore,
maintained 20 mol% of CdS relative to TiO2 as the optimum loading in subsequent studies with
the a-TiO2/Pt/CdS heterostructures. We must note that the role of the Pt co-catalyst in hydrogen
evolution has been well explored. A transmission electron microscopy image of a-TiO2/Pt/CdS
is shown in figure 6a. This image depicts uniform distribution of Pt and CdS nanoparticles on
the a-TiO2. The lattice fringes d spacing of 0.22 nm is assigned to the (111) plane of Pt.
Similarly, the d spacing of 0.35 nm is assigned to the (111) plane of the cubic CdS. Upon photo
irradiation, the charges generated in CdS are transferred to the interface of TiO2 and participate
in the water reduction. The activity in this kind of heterostructure, therefore, depends on the
surface/interface properties rather than the bulk properties of the TiO2. As hydrogen evolution
activity is considerably enhanced in the presence of benzyl alcohol–acetic acid as a sacrificial
agentix, we have used this mixture as the hole scavenger and carried out hydrogen evolution
reaction (HER) studies. We indeed found excellent improvement in hydrogen evolution,
yielding 31 mmol h−1 g−1 with AQY of 29% under visible light irradiation (figure 6b and table
2). As modification of the electronic structure of CdS by partial substitution of Zn as in
Cd1−xZnxS is reported to improve the visible light-induced HER activity in these
heterostructures, we have prepared heterostructures by partially substituting the Zn in CdS.
These heterostructures (a-TiO2/Pt/Cd0.8Zn0.2S) exhibited a maximum H2 evolution activity of
40 mmol h−1 g−1 with an AQY of 48% under visible light irradiation. Hence we assume the use
of amorphous TiO2 in such kinds of heterostructures has a potential to reduce the processing
cost and technologies.

Figure 15 (a) High-resolution transmission electron microscopy image of a-TiO2/Pt/CdS and (b) time course of H2 evolution
of a-TiO2/Pt/CdS as a function of time in the presence of benzyl alcohol–acetic acid.
The photocatalytic properties of ZnO/Pt/CdS heterostructures depend on the electronic, optical
and physical properties of ZnO. Conducting ZnO (F-doped ZnO) nanoparticles were therefore
obtained by the treatment of ZnO with NH4F (1 : 2 mole ratio) at 600°C in N2 atmosphere. We
have used F-doped ZnO for the purpose of forming ZnO/Pt/Cd1−xZnxS type heterostructures,
and used them in H2 evolution studies. Our preliminary results suggest that the H2 evolution
activity indeed increases with F doping in ZnO under visible light irradiation. The effect of
conductivity of ZnO is most remarkable in the absence of Pt. Detailed investigations will be
published elsewhere.

Narrow band gap of CdS (2.4 eV) relative to ZnO (3.2 eV) causes only CdS to get
excited upon visible light irradiation. As such ZnO does not participate in the light absorption.
Recently, we have developed a new approach to modify the electronic structure of wide band
gap oxides by co-substitution of N and F in place of O. For example, N, F co-substitution in
ZnO decreases substantially its band gap to 2.4 eV (figure a). Therefore, the structures
containing ZnO0.8(NF)0.2 and CdS provide two photosensitive centres. These structures
exhibited superior hydrogen evolution activities in the presence of both Na2S–Na2SO3 and
benzyl alcohol–acetic acid (figure b). Partial substitution of Zn in CdS in these heterostructures
also enhances the activity, as in the case of ZnO/Pt/Cd1−xZnxS.

Figure 16 (a) Comparison of UV–visible absorption spectrum of ZnO1−x(N,F)x with that of undoped ZnO (inset shows the
change in color from white to orange). (b) Comparison of effect of co-doping of N and F in ZnO on hydrogen evolution of
ZnO/Pt/CdS (25%) type hybrid nanostructures using Na2S–Na2SO3 or benzyl alcohol–acetic acid mixture as sacrificial agent
under visible light irradiation.
We have investigated the effect of the morphology and surface area of oxide
nanostructures in ZnO/Pt/CdS type heterostructures for visible light-induced hydrogen
generation. TiO2/Pt/Cd0.8Zn0.2S heterostructures containing TiO2 powder, TiO2 nanoparticles
(NPs), H2Ti3O7 nanotubes (NTs) and TiO2 NTs with surface areas of 40, 123, 248 and
249 m2 g−1, respectively, exhibited hydrogen evolution activity of 0.55, 1.01, 1.76 and
1.20 mmol h−1 g−1, respectively (figure a). Structures containing H2Ti3O7 NTs having highest
surface area show highest H2 evolution activity, whereas those containing TiO2 powder having
least surface area show least H2 evolution activity. We have also observed that ZnO/Pt/CdS
heterostructures containing (i) ZnO NPs, (ii) ZnO nanorod 1 (NR1), (iii) ZnO nanorod 2 (NR2)
and (iv) ZnO nanorod 3 (NR3) with surface areas of 45, 41, 18 and 9 m2 g−1, respectively,
exhibited hydrogen evolution activities of 5.36, 6.88, 5.29 and 2.55 mmol h−1 g−1, respectively
(figure b). Hence H2 evolution from these heterostructures follows the trend in the BET surface
area of the oxide nanostructures where higher surface area shows higher activity. A slight
variation from the trend is observed in the case of ZnO NR1 which showed higher
photocatalytic activity than ZnO NP. It is probably due to better separation of photo-generated
charge carriers in the case of ZnO NR1.

Figure 17 Photocatalytic H2 evolution of (a) TiO2/Pt/CdS and (b) ZnO/Pt/CdS heterostructures with oxides having different
surface areas.

Replacing expensive Pt with NiO in these heterostructures is a promising approach in


order to obtain cost-effective renewable energy production. NiO has several advantages over
Pt and other co-catalysts as it does not allow back reaction or recombination of H2 and O2. In
our further studies, we have replaced Pt with NiO in ZnO/Pt/Cd1−xZnxS heterostructures (figure
12). Although ZnO/NiO/CdS shows higher activities compared with ZnO/CdS, the most
remarkable improved activity is observed with partial substitution of Zn in CdS as in
ZnO/NiO/Cd0.8Zn0.2S. The activities obtained are comparable with those obtained with
ZnO/Pt/Cd0.8Zn0.2S and superior to those obtained with ZnO/Au/Cd0.8Zn0.2S. Similarly, we
have prepared heterostructures of TiO2/NiO/Cd0.8Zn0.2S by the deposition of Cd0.8Zn0.2S on
TiO2/NiO. TiO2/NiO was prepared by hydrolysis of aqueous solution of nickel acetate in a
solution containing TiO2 dispersion and subsequent thermal treatment at 400°C for 2 h in air.
We have used these structures for photocatalytic hydrogen evolution studies. Our preliminary
results show that use of NiO in TiO2/NiO/Cd0.8Zn0.2S results in H2 yields superior to those
obtained with TiO2/Pt/Cd0.8Zn0.2S with a wide range of molar ratios of TiO2 : NiO. Further
studies on determining the role of NiO in these heterostructures in separating charges and
catalyzing proton reduction are in progress. The detailed results will be published soon

elsewhere.

Figure 18 Comparison of photocatalytic hydrogen evolution with (a) different photo catalysts and (b) ZnO/X/Cd 0.8Zn0.2S
(X = Pt, NiO and Au).

There is considerable scope for developing heterostructures which are efficient, stable and
inexpensive for the reduction of water. More importantly, the structures should exhibit
H2 evolution activities at longer wavelengths (λ>600 nm). Detailed discussion on use of
heterojunctions or heterostructures for producing hydrogen is available in the literature.

2.2.2 Role of pH in Photocatalytic Water Reduction:

The extensive work directed to conversion of light energy to chemical energy’-“has met
with success in several recent schemes using tris (bipyridyl) ruthenium (II) to photo catalyze
water reductions. These schemes should be saw as model systems since they require a
sacrificial organic electron donor to drive the series of reactions leading to hydrogen
production. Recent work has indicated that these organic reagents may ultimately be replaced
by water, in the presence of an appropriate catalyst (e.g., Ru02).’ Since the completion of the
studies presented here, Griitzel has published a scheme using R~(bpy)~~+ as a photocatalyst
to split water into hydrogen and oxygen.~.~ One important aspect of these model systems is
their potential use as assay systems to evaluate the photocatalytic activity of metals other than
ruthenium. Use of these systems in this manner requires that they be well characterized. Kagan6
has reported a water photoreduction system including R~(bpy)~~+ as the photocatalyst, paraquat
(PQ2+) as the electron-transfer mediator (quencher), ethylenediaminetetraacetic acid (EDTA)
as the organic electron donor, and a colloidal platinum catalyst necessary for hydrogen
production. The proposed mechanism is shown in Scheme I. This catalysis is attenuated by the
competitive back-reaction R~(bpy),~+ + PQ+- - Ru(bpy),’+ +
PQ2+ The system is reported to
produce hydrogen over a pH range of 4-7. Although the general mechanism of this scheme has
been determined, little characterization has been done with regard to the quantitative effect of
solution pH on hydrogen quantum yield or to the role of the platinum ~catalyst.~ In particular,
the colloidal Pt cocatalyst has been suggested to function as a “microelectrode”9 providing a
surface at which the donor potential of the reduced quencher can be matched to the water
reduction potential, analogous to the potential matching at a bulk electrode. If this model is
correct, the total H2 yield might well depend on the potential of the photoreduced quencher
(and the pH). A second undetermined feature of these systems is the kinetics of reaction
between tris(bipyridyl)ruthenium(III) and EDTA, which in part determines the efficiency of
H2 formation. Therefore, we wish to report studies which clarify the roles of pH, mediator
potential, and EDTA oxidation rate in these water reduction systems.

2.2.2.1 Experimental work


Ru (bpy), C12.6H20 (G. Fredrick Smith Co.), paraquat (Aldrich), and Na2EDTA
(Sigma Chemical Co.) were used with no further purification.
Hydrogen quantum yields were measured on solutions containing a buffer (phosphate
or borate), Ru (bpy), C12 (2.0 X lo-‘M), Na2EDTA (2.0 X M), an electron-transfer mediator
such as paraquat (2.0 X M), and polyvinyl alcohol stabilized platinum catalyst.I0 The Pt
catalyst was prepared as described in ref 10, except that it was reduced by H2 gas for 2 h and
subsequently centrifuged at 150 OOO G for 1 h. A clear yellow solution results which contains
a highly active Pt catalyst. A similar preparation has recently been reported by GrBtzel? The
total concentration of Pt in the experimental runs was ca. lod M. The reactions were run at 23
OC under a nitrogen atmosphere. The stirred solutions were irradiated with a 300-W tungsten-
halogen lamp filtered to transmit light between 400 and 500 nm. Irradiation was continued until
the quantum yield became constant (4-6 h). Light intensities were measured with use of a
Reinecke’s salt actinometer.” Hydrogen was measured by gas chromatography with usage of a
Poropak Q column at 43 OC and nitrogen as the carrier gas.12 The electron-transfer mediators
listed in Table I were prepared by refluxing the desired ligand (G. Fredrick Smith Co.) in 1,3-
dibromopropane for several hours. The crude product was filtered from the solution, dissolved
in hot methanol/charcoal, and then reprecipitated by the addition of methyl ethyl ketone.”
Elemental analyses

Figure 19 Plot of electron-transfer quenching rate (k, = KSv/7) vs. electron-transfer driving force (Eo = E*Ruw2+ - E(Ru3+-
R,,z+ + Ed(Q-d

Confirmed the identity of the products. The El,* values (vs. SCE) of these mediators were
measured in aqueous 0.1 M KN03 solutions by differential-pulse polarography and cyclic
voltammetry with the use of a Princeton Applied Research Model 174A polarographic
analyzer. Stern-Volmer constants for the quenching of R~(bpy)~~+ luminescence by the
electron-transfer mediators were measured with a Perkin-Elmer MPF 44A fluorescence
spectrophotometer. Kinetics of Ru (bpy), + reduction by EDTA were monitored with a Dionex
D-1 10 stopped flow apparatus. Weakly acidic ruthenium solutions (lo-, M H+) were prepared
and used within minutes of preparation. EDTA solutions were buffered by addition of excess
acid (pH 2) or base (pH 4-6). A 10-fold excess of EDTA insured pseudo-first-order conditions.
Results and Discussion the pH dependence of the hydrogen quantum yield is shown in Figure.
In accord with previous results: an optimum pH of 6-7 was observed. Kagan has previously
suggested that the rate decrease above pH 7 occurs because the driving force for PQ+*
oxidation (EoOx = +0.4 V) is less than the reduction potential for water at pH 7 (EoRd = -0.45
V). In a more general form, this explanation suggests that water reduction (rate) is coupled to
the driving force for mediator oxidation. So that this general proposal could be tested, a series
of homologous electron-transfer quencher mediators were synthesized and characterized for
use in H2-generation system. A lit of these quenchers is contained in Table I. The reduction
potentials of these systems were characterized by cyclic voltammetry and differential-pulse
polarography on a hanging Hg-drop electrode. While PQ2+ and DMPB3+ showed good
electrochemical reversibility (AE(peak a - peak c) = 60 mV) the other mediators were only
quasi-reversible at lower scan rates. Hiinig and co-workers have previously reported
polarographic potentials for three of these systems.13 their values are also listed in Table I. We
believe that the discrepancies observed between the studies are largely due to the poor
electrochemical reversibility of several mediators. Given this irreversibility it is likely that the
fast scan CV measurements are more reliable than the previous polarographic data. The
bipyridyl-based species showed good electrochemical reversibility by CV. As expected, all
these mediators efficiently quenched the ruthenium excited state. Stern-Volmer quench
constants for these systems are also given in Table I. These values of K,, correspond to quench
rate constants of ca. 109 M-l/s-l. The quench behavior has been previously studied in detail for
PQ2+3 and more recently for PB2+.18c The relative quenching rates follow the mediator
potentials (Figure), as expected for an outer-sphere electron-transfer quenching reaction3 After
the electron-transfer properties of the mediators had been characterized, the effect of driving
force on H2 yield was studied. The dependence of the Hz quantum yield on electromotive force
is shown in Figure. A clear linear relationship between H2 quantum yields (aHz) (corrected for
relative quenching efficiency among different quenchers) and the driving force for water
reduction (EO) is observed. This relationship holds whether the driving force is altered by
changing the mediator potential or by changing the pH. This dependence is consistent with
control of H2 formation by “potential matching” at the Pt colloid surface, consistent with the
microelectrode model previously proposed9 and with Kagan’s explanation of the alkaline limit
for H2 formation. However, the control of (PHz is potentially complex, involving not only H2
evolution from the colloids but also the efficiency of separation of photoproduced redox
products. In the case of PQZ+, this efficiency is only ca. Thus, the observed (PH, dependence
might be complicated by pH or mediator dependent differences in initial redox product
formation. Factors which affect this efficiency include the rate of quenching (k,, obtained from
the Stern-Volmer constants in Table I), the rate of back reaction, kb, and the rate of reduction
of €tu3+ by EDTA, kRd. Differences in the quenching rates of different mediators have been
corrected for in Figure 3 by using the observed Ksy values in Table I to normalize to identical
quenching efficiencies. (Without this correction, significant deviations would be expected and
are observed). For any single quencher, varying the pH should not affect k, but would only
affect the driving force for water reduction. The back reaction rate in principle might depend
on the driving force for reaction and would increase as the quencher oxidation potential
increased. Thus, if kb changed for different mediators, this change would affect (PHz in the
opposite fashion to that observed. An increase in Eo of the mediator should increase kb and
decrease (PH,. Instead, as Eo increases, (PHz increases. This behavior can be rationalized by
the fact that the back reaction is essentially diffusion controlled for PQ+. + Ru3+ and will
proceed at a similar rate for the other mediators. (The self-exchange rates of the mediators,
estimated from the quenching rates with use of Marcus theory, vary less than a factor of 3.)
Thus although the back-reaction influences the overall quantum efficiency, it does not affect
the relative quantum efficiencies observed with the different mediators or pHs used here.
(Clearly pH should have little if any effect on kb.) In independent work, a0 it has been shown
that the same pH dependence of the rate of H2 formation is found for a photochemically
reduced mediator (as in the present work) as for an electrochemically reduced mediator. In the
electrochemical system, only an electrode (C or Hg), the mediator, and Pt colloid are present.
The fact that the same pH dependence of H2 rate is seen for the electrochemical system, in
which neither amine nor ruthenium is present, as for the photochemical system strongly
supports the above analysis that the H2 production rate is governed by potential matching at
the Pt surface. This control will only hold under conditions where the ancillary reactions
(excitation, quenching, and back-recombination) are essentially constant or corrected for, as in
the present case. Finally, the effect of pH on Ru3+ reduction by EDTA should be addressed.
(Changes in mediators should not affect this reaction). This reaction has been studied over a
range of pH as summarized in Figure 4. The rate of EDTA reduction decreases over 200 fold
between pH 7 (kRd) = 2 X lo6 M-I/sd) and pH 4 (kRd = 8 X lo3/ M-’/sd). This decrease likely
explains the previously observed.

Figure 20 pH dependence of the rate of reaction between R~(bpy),~+ and EDTA.

Cessation of H2 production below pH 4. In these sacrificial model systems, redox


product separation depends on competition between reduction of Ru3+ by EDTA and back-
reaction of Ru3+ with PQ+*. At pH potential varies from 0.13 V at pH 9 to 0.56 V at pH 2.
Over the pH range (6.5-9) used in the H2 yield studies, the EDTA reaction rate is already
maximized and should not change appreciably. The simplest explanation for the dependence
of a,, on Eo which is consistent with all the data obtained both by pH variation and mediator
replacement is that the H2 yield in the system reported here is determined by the matching of
the mediator redox potential to the water reduction potential at the Pt surface, analogous to
potential matching at an electrode surface. This data thus provides the first direct evidence for
the “microelectrode” model of the dispersed Pt ~catalyst.~J~ the work reported here, and
elsewhere, has sufficiently defined this “sacrificial” water reduction system that it may be
confidently used as an assay system to test the photocatalytic activity of metals other than
ruthenium. Indeed, with the use of such an assay, preliminary results have demonstrated
photocatalytic water reduction by chromium (III) 22 and metall~porphyrins. ~

2.2.3 Energy Requirements For Water Reduction

An essential problem to be solved is the coupling of the photoredox events with catalytic steps leading
to water decomposition. Here, we based our strategy on the idea of redox catalysis which was developed
in 1938 by Wagner and Traud [21]. To the solution are added finely dispersed catalytic particles which
serve as microelectrodes to give water oxidation or reduction selectively. Consider first the hydrogen
evolution step.

2R- + 2H2O + H2 + 20H- + 2R

In homogeneous solution this process is condensed difficult by the fact that is has to pass through the
stage of a free H' atom whose free energy is 2.1 eV above that of the H, molecule. In the presence of a
suitable catalyst the formation of free H' atoms is avoided which reduces considerably the energy
requirement for the water reduction step.

Figure 21 Pathways of light induced hydrogen evolution from water


The colloidal redox catalyst employed in our studies are functioning as local elements: the oxidation of
the reduced relay is coupled to hydrogen generation from water. The choice of the catalytic material
may be based on the same consideration which apply for electrocatalytic reagents used on
macroelectrodes: the exchange current densities for the anodic and cathodic electron transfer steps must
be high. Colloidal platinum would then appear to be a suitable candidate to mediate reaction. This fact
was recognized already at the end of last century when numerous examples for the intervention of finely
divided Pt in the process of water reduction by agents such as Cr2+ and Vz+ appeared in the German
colloid literature. A reaction of particular interest is the reduction of water by methylviologen.

2MV' + 2H20 - Hz + 20H- + 2MV2+.

The fact that this process can be catalyzed by Pt dispersions was discovered by Green and Stickland in
1934. More recently a number of photochemical systems have been developed in which MV2+ is
reduced in a light driven electron transfer reaction by a suitable sensitizer. The sensitizer cation
undergoes a subsequent reaction with a third component which is irreversibly oxidized. Such sacrificial
systems served to optimize conditions for the light induced hydrogen evolution. An illustrative example
is the case where Ru(bipy):+ serves as a sensitizer and EDTA as a sacrificial electron donor.

Figure 22 Correlation between the observed yields for Hz in the system 4. M Ru(bipy):+, M MV2+, 3. 10-2MEDTAatpH
and the radius of the catalysts employed. The lower points at 110 A contain 0.2 mg of Pt and 0.3 mg of Pt per 25 ml of
solution, respectively

Fig. 22 shows the influence of the radius of the Pt particle on the hydrogen evolution rates. These
particles were stabilized by polyvinyl alcohol. Illuminations were carried out with a XBO-450 W Xe-
lamp. One notices that a decrease of the particle radius from 500 to 100 A leads to a drastic augmentation
of the hydrogen evolution rate which is as high as 12 I/day/l solution for the smallest particle size. In
fact, in the last case, the bubbling of hydrogen gas occurring under illumination of the solution is readily
seen. In Fig. 10 the common feature is the Pt content of the solutions of different particle size namely
3.5 mg Pt/25 cc. This figure also shows two extra points with 100 A particle size but with low Pt levels;
0.35 mg Pt/25 cc solution, the upper one, and 0.25 mg Pt/25 cc solution, the lower one. Quantum yield
measurements indicate a stochiometric relation between the viologen reduced in the photoprocess and
the amount of hydrogen produced. One might object to the high Pt levels (140 mg/ltr) required to obtain
these high yields. Such concentrations are intolerable for practical systems. However, reduction in
particle size to a radius of only 15 A leads to a one hundred fold increase in the activity of the catalyst
[27]. The preparation of these ultrafine and stable Pt-particles is indicated schematically on Fig. 11. A
solution of hexachloroplatinate is reduced by citrate according to the method of Turkevich et al. [28].
Excess citrate is then removed by stirring with an exchange resin. After filtration an efficient protective
agent such as Carbowax-20 M is added or the particles are deposited on a powdered support such as
TiO2 or SrTi03. The very small Pt particle size obtained in this manner has been verified by photon
correlation spectroscopy and SEM. With such preparation a hydrogen output of - 10 ltr/day/ltr solution
can be achieved with a Pt concentration of only -1 mg/ltr.

Figure 23 Procedure for the preparation of ultrafine Pt-sols

A further advantage of these finely divided platinum dispersions is that the solutions remain completely
transparent even at high catalyst concentration. This allows to study directly the dynamics of the
reaction of MV+ with Pt particles by employing laser photolysis technique. The upper part of Fig. 12
shows oscilloscope traces illustrating the temporal behaviour of the characteristic MV absorbance at
602 nm in the absence and presence of catalyst. The upward deflection of the signal after the laser pulse
is due to the formation of MV+ via the photo redox process:

Ru(bipy):+ + MV2+ - Ru(bipy):+ + MV+

In this case, no back reaction occurs since Ru(bipy)!+ is reduced to the 2+ state by EDTA. While in the
absence of catalyst the MVt absorption is stable, addition of colloidal Pt induces a decay of the signal,
the rate of which increases sharply with Pt concentration. From a fitting of the absorbance decay curves
to an exponential time law, one obtains the rate constants which are plotted as a function of Pt
concentration in the lower part of Fig. 24. The ascent of the curve is steeper than linear, indicating that
the reaction order is greater than one with respect to the Pt concentration. At a concentration of M Pt in
Pt-PVA-polymer, we observe a rate constant k = 0.14 - lo' sec-'. At the highest concentration of catalyst
(12.5. M Pt content), the rate was 5.7. 104 sec-'. The lifetimes observed for MV+ were shortened from
about 7000 ps to 15 psec, when the concentration of catalyst vaned only by a factor of about 12. The
high reaction rate of the PVA protected platinum catalyst (RH = 110 A) with the reduced viologen is
even exceeded by the rates obtained with the 15 A radius Pt particles. In this case the half lifetime of
MV2+ in the presence of M Pt (20 mg/l) is only ca. 30 ps*). Taking into account that one particle
contains 1200 Pt atoms, one derives a rate constant for reaction (7) of ca. 2.10'' M-' s-' indicating that
this process is essentially diffusion controlled. These kinetic results are of great importance for the
design of a cyclic water decomposition system. It appears that, by suitable choice of the catalyst, the
conversion of A- into A and simultaneous formation of hydrogen can be accomplished so rapidly that
it can compete efficiently.

*) This is the limiting rate obtained at pH < 1 with the thermal back reaction. (In the photo stationary
state achieved under sunlight irradiation, the latter process, due to its bimolecular nature, should require
at least several milli-seconds).
Figure 24 Mechanism of intervention of the ultrafine Pt particles in the H2-formation, electron pool effect

The particle is polarized cathodically through electron transfer from the reduced relay species. Coupled
with this process is the hydrogen ion discharge leading to the formation of adsorbed hydrogen atoms.
The latter subsequently combine to yield H2. The concept of coupling of the two electron transfer
processes is supported by the fact that the rate of disappearence of R- is acid catalyzed. In the case of R
= MV2+, this fact has been clearly established by using pulse radiolysis technique.(Grätzel, 1980)

2.2.4 Bifunctional water splitting catalyst

The case of the water oxidation by the Ru(bipy)i+ complex is particularly interesting as this species is
formed in the photo reaction which preceeds hydrogen generation from the reduced relay. One
recognizes here the possibility to close the cycle of water decomposition by visible light. In fact,
irradiation of a solution containing colloidal Pt (protected by a copolymer of maleic anhydride and
styrene) and macrodisperse RuO, as catalysts, apart from the Ru(bipy)$ sensitizer and the MV2+
electron relay, produces the two gases hydrogen and oxygen simultaneously [33]. The successful
operation of this system depends on the proper choice of the protective agent for the Pt sol. The
copolymer of maleic anhydride and styrene employed in our experiments is suitable in that it provides
functions with pronounced hydrophobicity. Of the redox products formed in the light reaction,
Ru(bipy)i ' is strongly hydrophylic while MV+ is relatively hydrophobic. Hence, the latter will
predominantly interact with the Pt particles giving rise to hydrogen formation. The Ru(bipy):+ left
behind is prone to interact with the hydrophilic RuO, surface yielding oxygen from water. In this system
the reaction rate of MV+ with the Pt particles must also be high enough to be competitive with back
electron transfer and the reduction of oxygen.

MV+ + 0, - 0, + MV2+.

Figure 25 Scheme for cyclic water/decomposition in a combined catalytic system

The overall reaction scheme is illustrated in Fig. 25. With the combination of catalysts employed
initially the quantum yield of water splitting was found to be relatively low, i. e. 1.5. However, this
figure has, meanwhile, been improved drastically through the utilization of a bifunctional redox catalyst
[34]. Here, Ti0, particles doped with Ru02 were employed which serve at the same time as support for
an extremely fine Pt deposit. Surprisingly, with such dispersions one achieves already 1/5 of the
quantum yield obtained with the sacrificial systems. Significantly, the Hz and 0, production rates do not
decrease over many hours of irradiation time. This is illustrated in Fig. 26 which shows the amount of
hydrogen produced by illuminating a solution containing 10-4 M Ru(bipy)2/3 +, 2 . 10-3 M MV2+ and
the bifunctional catalyst. The hydrogen formation rate remains constants over at least 40 hours of
irradiation time, oxygen being produced in stochiometric proportion. The quantum yield of hydrogen
formation in this domain is ca. 2%. This shows that the extent of cross reactions is less important with
this bifunctional catalyst.
Sustained light induced water splitting in a system containing bifunctional redox catalyst, (Ti02/RuO)

Figure 26 Sustained light induced water splitting in a system containing bifunctional redox catalyst, (Ti02/RuOz/Pt)

It is likely that adsorption of the reactants and/or participation of electronic states of the TiOz
semiconductor in the redox events render the water splitting process so efficient. Fig. 19 illustrates a
feasible mechanism for the intervention of the semiconducting redox catalyst. Light excitation of the
sensitizer (Ru(bipy):+) is followed by electron transfer to the relay - in our case methylviologen. The
latter in turn injects the electron into the conduction band of the semiconductor. From there on it is
channeled to a Pt-site where H2 evolution occurs. Ru02 on the other hand assists the back-conversion
of S+ to S under simultaneous oxygen formation from water. Ti02 is a favorable material in that its
conduction band is located close to the H2/H+-standard potential. The chemisorption of O2 to this
material may further assist the water cleavage by light. An alternative approach to achieve light induced
cleavage of water is to separate the hydrogen from the oxygen generating reaction. A device which
achieves this goal is depicted schematically in Fig. 20. The anode compartment contains simply water
and a Ru02 electrode and is kept in the dark. It is coupled via an external circuit and a proton conducting
membrane to the cathode half-cell which is illuminated. The latter contains a sensitizer, an electron
acceptor (R) and a platinum gauze electrode. Light quanta drive the electron transfer from S to R
producing the species S+ and R-. The latter in the presence of a suitable catalyst will reduce water to
hydrogen. The back conversion of the oxidized sensitizer into its original form occurs via the external
circuit. Electrons are furnished from water oxidation occurring at the Ru02-electrode.
Figure 27 Cell device for cyclic water decomposition, oxygen and hydrogen evolution occurs in two separate half cells

At present, we have concluded a study on a photoredox system where Ru(bipy)2/3+ was employed as a
sensitizer and peroxodisulfate as an acceptor. The latter is irreversibly reduced to sulfate in the
photoreaction. Photopotentials obtained with such a system are sufficiently high to afford water
oxidation even under room light. The photocurrents are of the order of several hundred µA/cm2.
Furthermore, a catalytic system is presently explored where the photoredox reaction is followed by
reduction of water making the process cyclic with respect to the acceptor. Such a cell system should
also afford water cleavage by visible light with the additional advantage of producing H2 and O2
separately.(Rao et al., 2016)
References

BARENDRECHT, E. 1982. Electrochemical energy storage and conversion. Pt-Procestechniek, 37,


52-57.
GRÄTZEL, M. 1980. Photochemical Methods for the Conversion of Light into Chemical Energy.
Berichte der Bunsengesellschaft für physikalische Chemie, 84, 981-991.
MAEDA, K. & DOMEN, K. 2010. Photocatalytic Water Splitting: Recent Progress and Future
Challenges. The Journal of Physical Chemistry Letters, 1, 2655-2661.
MEZA-CHINCHA, A.-L., SCHINDLER, D., NATALI, M. & WÜRTHNER, F. 2021. Effects of
Photosensitizers and Reaction Media on Light-Driven Water Oxidation with Trinuclear
Ruthenium Macrocycles. ChemPhotoChem, 5, 173-183.
NAJAFI, L., BELLANI, S., OROPESA-NUÑEZ, R., PRATO, M., MARTÍN-GARCÍA, B., BRESCIA,
R. & BONACCORSO, F. 2019. Carbon Nanotube-Supported MoSe2 Holey Flake:Mo2C Ball
Hybrids for Bifunctional pH-Universal Water Splitting. ACS Nano, 13, 3162-3176.
RAO, C. N. R., LINGAMPALLI, S. R., DEY, S. & ROY, A. 2016. Solar photochemical and
thermochemical splitting of water. Philosophical Transactions of the Royal Society A:
Mathematical, Physical and Engineering Sciences, 374, 20150088.
SCHNEIDER, J. & BAHNEMANN, D. W. 2013. Undesired Role of Sacrificial Reagents in
Photocatalysis. The Journal of Physical Chemistry Letters, 4, 3479-3483.
TEE, S. Y., WIN, K. Y., TEO, W. S., KOH, L.-D., LIU, S., TENG, C. P. & HAN, M.-Y. 2017. Recent
Progress in Energy-Driven Water Splitting. Advanced Science, 4, 1600337.
WUBBELS, G. G. 1983. Catalysis of photochemical reactions. Accounts of Chemical Research, 16,
285-292.
YOUNGBLOOD, W. J., LEE, S.-H. A., MAEDA, K. & MALLOUK, T. E. 2009. Visible Light Water
Splitting Using Dye-Sensitized Oxide Semiconductors. Accounts of Chemical Research, 42,
1966-1973.
ZHAO, C., YANG, X., HAN, C. & XU, J. 2020. Sacrificial Agent-Free Photocatalytic Oxygen
Evolution from Water Splitting over Ag3PO4/MXene Hybrids. Solar RRL, 4, 1900434.

You might also like