You are on page 1of 12

The major issue to be handled in the near future is the enormous energy needs.

Researchers are increasingly committed to researching clean, safe, and sustainable energy
options to address anticipated nonrenewable energy shortages and monitor environmental
pollution. Hydrogen gas (H2) is considered a sophisticated source of renewable energy. It
is probable to replace conventional fossil fuels with sustainable H2 fuel. It exhibits a high
value of gravimetric energy in comparison to that of the gasoline (i.e., 120 MJ/kg versus
44 MJ/kg). Its nature is storable and renewable, and no carbon emissions occur from its
consumption. The H2 fuel can be generated through various means, including steam
reforming/partial hydrocarbon oxidation, coal gasification, and water splitting. The
Earth’s surface is mainly covered with 71% of water, and the purest form of H2 is
produced from water splitting, which does not require high-temperature treatment. Thus,
in order to reduce the anthropogenic CO2 emission in the atmosphere, solar water
splitting has been an essential factor. The need for the development of scalable, practical
and clean energy capture, generation and storage systems has spurred vast amounts of
research in the recent years. Solar energy is ubiquitous to the clean-energy discussion
given its scale (~120 000 TW average irradiation at the earth's surface). Despite it being
the largest energy source of the planet, direct solar-energy utilization only accounts for
less than 0.06% of the global electricity generation. Economic and implementation
challenges are the most important causes for the low dissemination of solar-driven energy
generation systems. The price of photovoltaic (PV) modules has declined significantly in
the past decade (5–7% annually) leading to a continued increase in their deployment and
grid integration. However, this energy is of intermitted nature and adds complexity in
balancing the grid load. The challenge of efficiently using intermittent sources of energy
has resulted in significant interest towards the development of economically viable and
scalable energy storage solutions. Currently, energy storage takes place in its vast
majority via pumped-hydroelectrical systems (more than 99% of storage, with a total 127
GW capacity). Although this solution is economically viable in certain instances, its
implementation is constrained by geographical factors and can only serve as a mean of
central energy storage with limited usability in the transportation sector. Less wide-
spread technologies for energy storage include compressed air, flywheel systems, thermal
storage systems and batteries, some of which can be used for both stationary and mobile
energy generation.
Our emancipation from fossil fuels in the long term will have to be supported by
solar energy capture. The average global energy consumption rate today is 15 TW and
estimated to increase to 30 TW by 2050. Renewable sources such as wind and biomass
could each deliver less than 10 TW when deployed all over the world. The potential of
solar energy however seems practically unlimited. Fossil fuels, plants and wind are all
derived from solar energy. H2 is central to the development of a hydrogen economy,
which is already starting today. About 100 H2 fueling stations exist in the United States,
most of which are in private ownership by companies. Los Angeles hosts 8 public fueling
stations, and 100 more are planned in California by 2024. In Europe, Germany is at the
forefront of public H2 fueling stations. The current number of ca. 20 is to be augmented to
400 by 2023. A solar-driven station was planted at Freiburg, which produces its own H 2
using solar energy captured by PV panels on the roof. Almost 20 such stations exist in
Germany and the US. A solar H2 fueling station for residential use was presented by
Honda, and Nanoptek developed a water splitting device that combines an integrated
photocatalyst with external PV panels. One of the main energy applications of H 2 today
remains its use in forklifts in distribution centers. Over 6 200 such forklifts are in
operation in the US, used by companies as Walmart, Sysco, Central Grocers and BMW.
Colruyt, a Belgian retailer, together with WaterstofNet and Hydrogenics installed a H2
fueling station that is powered by their own excess wind and solar energy. Another
success story of H2 was found in public transportation, with dozens of H 2 based buses
involved in pilot projects in cities such as London, Beijing, Berlin and Perth. Despite the
deployment of H2 infrastructure on several locations, solar production of H 2 remains a
challenge. The main factor to deal with is cost, illustrated by the change of strategy of
Sun Catalytix, a spin-off company of the Massachusetts Institute of Technology. It aimed
to commercialize its ‘artificial leaf’, a solar device that directly produces H 2 from water.
Despite its simple design and use of earth-abundant materials, it could hardly compete
with conventional PV. Widespread solar H2 is within reach, and the final steps toward
practical application are outlined in this paper. We will discuss two approaches:
 a conventional water electrolyzer driven by an external PV panel (PV/electrolysis).
 a Photo Electro Chemical (PEC) cell that integrates photoactive and electrochemical
components in a single device.
A fundamental solution to ever-growing concerns on global warming, air
pollution, and energy security is to replace the current carbon-rich fossil fuels with
renewable and environment-friendly carbon-free or carbon-neutral energy sources. Fig. 1
provides a brief view of the energy reserves available on the earth, which demonstrates
the dominant position of solar energy among all renewable and non-renewable energy
sources. The total sum of recoverable energy from all the reserves is around 1% of the
solar energy supplied to the earth surface by the sun. Moreover, the solar energy (173 000
TW) supplied to the total surface (land and water) of the earth is nearly 9600 times larger
than today’s total energy consumption of the world (17.91 TW in 2017). Evidently, solar
energy is much more than enough to meet all energy needs for the lifetime of the solar
system, but the challenge lies in the development of economically viable technologies for
its harnessing, storage, and utilization. There are several technologies available for solar
energy utilization – solar thermal (Fig. 1c), photovoltaic (PV) (Fig. 1d and e), and solar
fuels (Fig. 1f and g). Today, PV represents the most widely disseminated technology and
is one of the fastest growing industries. But PV produces electricity, which has an
intrinsic difficulty in terms of storage. Considering the intermittent nature of solar energy,
converting solar energy into easily storable chemical energy or a fuel has emerged as an
ideal option of solar energy utilization. Among the potential solar fuels, hydrogen (H 2)
generated from solar water splitting (Fig. 1g) has received attention as the most
promising fuel especially for powering fuel cell vehicles. ith the progressive research
efforts in developing efficient photocatalytic system for durable solar water splitting, it is
timely to deliberate on scaling up of this technology for practical application. As
delineated in figure 42, it can be clearly visualized that both PV-EC and PEC confer
higher complexity than photocatalysis system which indicates higher cost and higher
solar efficiency target for practical application. As mentioned earlier, the PEC cell is
located between PC and PV–EC in its complexity, choice of material (light absorber),
and obtained overall efficiency.
Figure 42
Photoelectrochemical (PEC) water splitting offers a sparkling and sustainable
strategy for hydrogen generation. The PEC water splitting process has been regarded as a
promising approach for solar energy conversion to H2 fuel. This process offers a low
input bias to offset the overpotential for water oxidation/reduction. Water oxidation (i.e.,
oxygen evolution reaction (OER)) or water reduction (i.e., hydrogen evolution reaction
(HER)) could be achieved via the PEC water splitting at potentials below 1.23 V and
above 0 V compared to the reversible hydrogen electrode (RHE), respectively. These are
the potential steadiness values for two half-cell reaction processes at room temperature,
respectively. The PEC approach is preferable over the other water splitting techniques
due to numerous motives described below =
1. First, to initiate the PEC water splitting reaction, the efficiency of solar-to-H 2 (STH)
is better, whereas many solar photovoltaic devices (e.g., solar cells) are required for
providing the appropriate voltage for electrocatalytic water splitting.
2. Second, during PEC water splitting, high photocatalytic activity for OER and HER
cannot readily be accomplished simultaneously, while a high level of OER and HER
activities may be obtained via the sensible selection of PEC water splitting
photocathodes and photoanodes.
3. Third, pure H2 and O2 products are readily separated and produced by the PEC, and
photocatalytic water splitting critically affects the environmentally friendly separation
of O2 from explosive H2 and O2 combinations.
4. Fourth, high temperature and numerous steps are not involved during the PEC water
splitting process.
Sustainable solar H2 production systems are based on three water splitting
systems1 as schematically depicted in Fig. 4; (i) particulate PC systems, (ii) PEC water
splitting cells, and (iii) PV–EC systems. In a particulate PC system, the photocatalyst
powders are dispersed in the medium, and thus the charge transfer pathway is
significantly shorter (several micrometers) than usual charge collectors in the other
systems (sometimes tens of meters). Since both H2 and O2 are generated from each of the
PC particles, a separate gas separator is needed. The system is the simplest with only a
light absorber and an electrocatalyst. But when the Z-scheme is used, the system becomes
D4 and needs a charge mediator adding complexity. In a PV–EC system, independent
power is generated by a PV device, which is connected to electrocatalysts. Thus, the
performances of PV and electrocatalysts are of primary importance, while charge transfer
between them does not matter. The light absorber is completely out of water, and hence,
stability against corrosion in the aqueous medium is of little concern. Also, sizes of PV
and EC components can be freely modulated because they have very different balance of
system (BOS). The BOS denotes auxiliary components needed for operating the device
such as piping, controller, fuel or charge transfer roads, which will be depicted in more
detail later. A disadvantage of the PV–EC system is the highest complexity in PV
designs, but most of the related technologies are already mature and commercialized. In a
PEC cell, photoelectrodes are connected to charge collectors, and only the light absorber
side of the photoelectrode contacts the liquid medium while the charge collector is
isolated. Fuel separation can be achieved spatially because H 2 and O2 evolve from
different electrodes. Band bending is quite obvious relative to PC along the photoanode
and photocathode for facile charge separation and transfer. A unique semiconductor–
liquid junction (SLJ) is established as described above, but it is also possible to use a
semiconductor with a buried junction. For these cases of semiconductor–semiconductor
junction (SSJ) (e.g. Pt/TiO2/p/n-Si), metal–insulator–semiconductor (MIS) (e.g.
Pt/SiO2/n-Si), or electrocatalyst–metal oxide–semiconductor (e.g. Ni/SnO2/p/n-Si as
photoanode), their driving force of charge transfer comes from the built-in potential from
the interface of solid-state materials, not from the semiconductor-electrolyte interface.
In this section, we graphically summarize in Fig. 21 the performance of various
solar water splitting systems discussed so far in terms of STH conversion efficiency
ZSTH and stability for three system categories of PV–EC, PEC, and PC based on data
summarized in Tables 1–3. Similar summaries were presented in 2014 by Ronge et al.
and in 2015 by Ager et al. In addition, we subdivide PEC devices into PE buried (devices
similar to PV cells but function in water with sophisticated surface protection), PE–PV
(at least one side of the device directly contacts water forming a semiconductor||liquid
junction, SLJ), and PE–PE (made with only photoelectrodes) devices, all having a main
light absorber working inside of the electrolyte. In contrast, PV–EC devices work outside
of the electrolyte, and PC devices are free or fixed powders with no charge collector or
photoelectrodes. In general, PV–EC systems clearly show the highest ZSTH (up to 20%
by III/V, B10% by Si), followed by PEC (3.7% for PE–PE, B8.1% for PE–PV, B19% for
PE buried), and then PC (mostly around 1.0%) systems. Also, there is an appreciable
trend of steady improvement in ZSTH over the last two decades. The PEC systems show
the largest variation of ZSTH because of the largest system variation. Thus, the systems
of PE buried show the most complexity and the highest ZSTH, while PE–PE systems of
the minimum complexity show the lowest ZSTH. Unfortunately, none of those
experimentally studied systems meet the criteria for practical solar hydrogen production,
i.e., a high efficiency above 10%, a long lifetime over 10 years, and scalability up to
several km2, which we discuss in more detail in later parts of this paper. Benchmarking
stability is not simple because the reported period of stability does not necessarily mean
the lifetime of the device. But the trend is clearly seen. The best stability is observed with
PC or PEC systems using intrinsically stable oxide semiconductors or C 3N4. However,
there are few p-type semiconductors that are stable under photocathode conditions. Many
successful cases of protection layer strategies have been reported for unstable p-type or
PV grade materials. However, the practical devices require durability as long as 10 years,
and longevity of the protection layers for such a time scale is of concern. There is a
review specifically addressing the stability issue of photoanodes. It also should be
mentioned at this point that efficiency and durability are only two of the many factors
that determine the economic competitiveness of the system as will be discussed in later
sections.

To capture 30 TW of solar energy using 10% efficient Photovoltaic (PV) panels,


0.24% of the Earth’s surface is needed. Although this may seem a small number in terms
of land usage, there are serious technical challenges. To put this
number into perspective, we note that currently ca. 0.4% of the Earth’s surface is covered
by roads in rural areas, and ca. 2.8% of the Earth’s surface is urbanized. The main hurdle
impeding large-scale implementation of solar energy is its unreliability. PV produce
electricity, which is difficult and costly to store. Due to the intermittent nature of solar
energy, it is currently difficult to integrate large capacities of PV into an energy supply
strategy. For example, Germany drastically reduced its dependence on nuclear energy
and is investing strongly in renewables, but now relies on coal and gas fired power plants
to balance its energy supply and demand. Additionally, models predict an overload of
electricity grids when operated at high renewables capacity, invigorating the need to
buffer energy generation using energy storage technologies.
REQUIREMENTS FOR PRACTICAL SOLAR HYDROGEN SYSTEMS
1. Cost
Most of the H2 today is produced via steam methane reforming, at prices down to 1 $ kg -1
H2. Competing technologies are centralized biomass gasification and wind electrolysis,
with prices of 1.6 $ kg-1 and 4.5 $ kg-1, respectively (distribution costs not included).
Solar H2 is not yet cost effective today. In its Hyways roadmap, the European
Commission targets a distributed solar H2 price of 4 £ kg-1 by 2020 and 3 £ kg-1 by 2030.
The US Department of Energy aims for 5 $ kg -1 by 2017 and 4 $ kg-1 by 2020
(distribution costs not included). A distributed H2 price of 2-4 $ kg-1 is assumed to be cost
competitive with H2 derived from fossil fuels. Parkinson and Turner argue that this metric
is unrealistic as it directly compares the cost of renewable and methane derived H 2.
Instead, they propose to compare costs based on affordability, factoring in also the cost of
greenhouse gases emitted by fossil fuels. Taking gasoline as a benchmark, they propose a
target H2 cost of 6 $ kg-1. A technoeconomic analysis of centralized facilities based on
PEC cells predicted that prices between 2.9-18.8 $ kg -1 could be achieved, including
pressurization of the H2. The analysis also considered costs related to land usage,
personnel, maintenance and process control. The large variation in cost estimation is due
to a number of uncertainties:
 capital cost,
 efficiency,
 lifetime.
The influence of these parameters on a PEC cell is given in Figure 4. Several scenarios fit
the targeted H2 cost. A cheap and efficient device is allowed to have a shorter lifetime,
although there are practical limitations and costs increase exponentially at lifetimes of < 5
years. Low-cost PEC cells with efficiencies around 5% have been demonstrated, but in
the long term an efficiency of 15% would substantially facilitate reaching the cost
objectives. The cost could also be reduced under the influence of a carbon tax or
government subsidies. On the other hand, PV panels and electrolyzers have been
investigated for decades and only incremental improvements to their efficiency are to be
expected in the future. Capital cost and electrolyzer lifetime are the most important
parameters to bring down the cost of H2 from PV/electrolysis. Whereas the cost of PV
panels has been steadily decreasing for decades, it may prove difficult to sufficiently
lower the cost of existing electrolyzers employing expensive polymer membranes and
noble metal catalysts. Although a comparative assessment of both solar H 2 approaches
was not yet performed, PEC cells bear the promise to reduce capital costs since all
components are contained in a single system, and possibly some components can assume
multiple functions (e.g. PV and membrane functions can be integrated in a single
material). Figure 5 shows the electricity price from a PV/electrolyzer/FC system as a
function of PV electricity price, without considering electrolyzer capital costs or storage
and distribution costs. For a corresponding H2 price of 2-4 £ kg-1, PV electricity should be
delivered at 0.05-0.1 £ kWh-1. In reality lower prices are desired, as additional capital and
distribution costs have to be covered. H2 cost has been assessed for centralized H2
production via integrated PV/electrolysis plants. Without including storage and
distribution costs, with 10-14% efficient PV and a 20-30-year lifetime H2 cost was
estimated at 4.70- 5.86 $ kg-1. If PV plants would reach lifetimes of 60 years and 16%
efficiency, H2 prices could go down to 1.90 $ kg-1.
2. Lifetime
The influence of lifetime on cost is apparent from Figure 4. PEC cell lifetime is
determined mainly by corrosion of semiconductors, catalyst poisoning and membrane
contamination by salts. Many of the state-of-the-art PEC cells operate in highly acidic or
basic electrolytes which more easily cause corrosion of semiconductors and catalysts.
Working in near-neutral electrolytes would relax the stability requirements but problems
arise due to the depletion of supporting electrolyte. A PEC device operated in near
neutral medium has been proposed with recirculation of the electrolyte to maintain
stability over longer periods. A lot of work has been done recently on protective ultrathin
coatings to prevent corrosion. Many PEC cells are found to be stable at least for several
hours. The US Department of Energy put forward a target of 5 000 hours stable operation
by 2018. In PV/electrolysis cells, the PV component is not exposed to a harsh
environment and will probably last as long as conventional PV. In this case, the
electrolyzer may be replaced when necessary while the PV remains. Performance
stability over time is also important. If the efficiency degrades, this again badly affects
H2 cost. On the other hand, the amount of maintenance required has an influence on cost,
capacity factor (downtime during maintenance) and practical applicability. High-
maintenance systems are less appropriate for remote or residential applications.
3. Scalability
When increasing the scale of solar H2 systems from lab to industrial scale the following
issues should be dealt with:
 engineering considerations
 materials and fabrication procedures
 safety and environmental issues
Engineering Consideration
Reactor engineering is an aspect which is seldom discussed in literature, but which may
have a large influence on the choice of candidate materials. Reactors should be as simple
as possible, low cost and low maintenance. PEC cells have the advantage of device
simplicity, as no separate electrolyzer is required. On the other hand, the integration of
multiple functions in a single reactor may complicate the interior architecture of the
device. An increase of a panel’s surface area should result in a proportional increase of
the H2 output. Current lab-scale systems often do not consider mass transport energy
losses when device dimensions are increased. Haussener et al. showed by numerical
modeling that for devices based on dense PV, increasing surface area may result in
tremendous overpotential losses. This was ascribed to ionic transport difficulties. On the
other hand, electronic transport can be limiting when electrodes are based on transparent
conducting oxides such as indium- or fluorine-doped tin oxide. Such back contacts have
low conductivity and fail even at dimensions of a few cm2.
Fabrication and Materials
If solar H2 generation is to be deployed on a global scale, the use of abundant materials is
imperative. In this respect widespread use of electrolyzers containing high loadings of
scarce Pt seems unlikely. Electrolyzers and PEC devices can circumvent this
contradiction either by drastically reducing Pt loading or by using earth-abundant
elements (Co, Ni,Fe). In either case, the device will have to be run at lower current
density. PEC cells have low current density limited by solar illumination, and are
perfectly suited to incorporate catalysts made of abundant elements. For PV/electrolyzer
setups, the use of less noble metals will require an increase in electrolyzer surface area.
Many PEC cells contain metal oxides based on earth-abundant elements (BiVO 4, WO3,
Fe2O3, Cu2O), but the most active examples also contain PV junctions based on Si. Even
though Si is the second most abundant element in the Earth’s crust, the requirement of
very high purity Si for PV devices makes their production process costly. Thin-film
technology (amorphous silicon) is gaining a lot of interest as it requires bless material,
although efficiency generally is also lower. Besides the choice of materials, also the
fabrication method should be suitable for scale-up. Often plasma-assisted vapor
deposition methods are used which require vacuum pumping, but milder wet methods
such as electrochemical deposition or even inkjet printing exist as well.
Safety and Environmental Issues
Solar H2 installations only make sense if they generate more renewable energy than it
costs to fabricate them. This has been investigated by Life Cycle Assessment (LCA).
LCA for large scale industrial PV/electrolysis has been carried out by Mason and
Zweibel, partly based on data from existing PV plants. They considered PV efficiencies
of 10-14% and a lifetime of 20-60 years. They show that 50% of the primary energy
requirement is for the PV power plant, 9% is used in the electrolysis plant and the
remaining energy is used for compressing and transporting the hydrogen fuel to fuel
stations. For PEC cells, an LCA has been performed by Zhai et al. This study has a much
wider range of uncertainty as the technology of PEC devices is not yet ready for practical
applications. Three cases were considered with different types of materials, PEC
efficiencies of 3-10% and lifetimes of 5-30 years. They showed that photoelectrode
fabrication is the most energy-consuming process. The results indicate that the energy
needed to build PEC plants could be significantly lower compared to PV/ electrolysis
(Fig. 6). A 10% efficient PEC device with a lifetime of 30 years requires as little as 10
MJ kg-1 H2 primary energy, whereas a PV/electrolysis plant with 12% efficient PV and 30
years lifetime is estimated to require 23.7 MJ kg -1 H2. Energy cost of H2 pressurization
and distribution is not included in these numbers, but should be comparable for both
approaches. This analysis shows that even though PEC devices capture less solar energy
per m2 (Fig. 2), primary energy requirement is lower due to a lower capital energy cost.
Even though solar H2 generation is still under intensive development, some of the
existing technologies can and are already being commercialized. Besides continued
research, commercial development should already begin today, as reaching market
penetration may be an even bigger challenge than the technical challenges solved in the
past. This article reviewed the most important practical considerations for PV/electrolysis
and PEC systems. Both deserve attention, as it is not yet clear which approach will be the
most successful in the future. While PEC cells hold the promise of lower cost, lower
environmental impact and less complexity, at the moment reality is arguing against this
technology, which is still facing many technical difficulties. Will PEC cells become the
next nuclear fusion? Or will they soon overcome their limitations and become the next
big renewable technology? We should hope for the latter, and in the meantime make use
of the possibilities PV/electrolysis give us. PV/electrolysis systems are currently the most
mature technology and more likely to be cost-effective on short term. As PEC devices
mature, more detailed technoeconomic and life cycle assessments are desired. Only then
will we be able to determine whether they are a useful technology. In the end, probably
both systems will be used depending on the application. The world’s future energy
demand of 30 TW by 2050 will have to be met by a balanced mix of wind, solar and
other sustainable technologies, deployed at large to small scales, and backed up by
reliable storage media. A very simple PEC cell design is possible if water vapor from air
is used rather than a purified liquid water source. In this case, the PEC cell requires no
input but sunlight and air. We provided a first assessment of such a concept by numerical
modeling. We showed that supply of water vapor by natural convection and wind is not a
problem. However, optimal performance at reduced relative humidity will require
appropriate reactor and materials engineering.
Although these processes have the common goal of splitting water into hydrogen
and oxygen, the economics of each technology is influenced by distinctly different
factors. PV-electrolysis depends on developments in photovoltaic cells - both price and
efficiency. Photocatalytic and PEC water splitting depend on the reactor design and
material development. Although this assessment considers a well-to-gate basis, it is worth
pointing out that formidable post-processing challenges still remain for solar generated
fuels such as hydrogen. Notable hurdles include the storage and transportation of
hydrogen as well as safe and commercially viable utilization in fuel cells. Work in these
areas is ongoing with incremental improvements driving progress.
3.1 PV-electrolysis of Water
The economics of PV-electrolysis is very similar to that of grid electrolysis.
However because the electricity is generated 'in house' with a PV panel, the economics
are not as strongly correlated to the grid electricity price. Economic feasibility is a
function of the performance of the system (PV cells and electrolysis units together). This
means an improvement in the PV technology is a critical factor and commercial success
relies upon development of the technologies performance. As electrolyzers are relatively
mature at current stage this discussion will focus primarily on the potential advances in
photovoltaic cells.
3.1.1 Technology Drivers
The fundamental difference between electrolysis and PV-electrolysis is the source
of the electricity feedstock. The former uses grid electricity whist the latter obtains
electricity from photovoltaic cells. One of the most important reasons for using
photovoltaics is the environmentally responsible and sustainable sourcing of electricity
feedstock. Recent awareness of environmental issues and a resource constrained future
has certainly pushed energy company sentiment in the direction of PV. However for
many, unless economical, this alone is not enough to incentivise the use of PV
technologies. PV must be an economically competitive technology which requires
continued performance development. Incremental improvements and learning curves in
1st generation (silicon wafer) solar cells have gradually reduced the costs of PV
technology. Although efficient, 1st generation solar cells are expensive because they
require thick wafers and vacuum processes for film fabrication. Despite this they are
expected to remain the dominant PV technology into 2020 when 2nd generation systems
will become prevalent. The 2nd generation (thin film) solar cells are fabricated by
deposition of photosensitive materials such as silicon, Cadmium Telluride (CdTe) and
Copper-Indium-Diselenide (CIS). Thin films, although at the lower end of the PV
technology efficiency spectrum, offer a low-cost option which suits large scale
applications where land cost is not significant. The long term future of PV devices may
however be in 3rd generation solar cells which seek to combine the advantageous aspects
of the 1st and 2nd generation technologies. Promising methods encapsulated in the 3rd
generation umbrella term include multi-junction cells, intermediate-band cells, hot carrier
cells and spectrum conversion. Taken from literature, Table 3.1 gives a technical
summary of PV generations and absorption materials. What is essential here is that vast
amounts of potential and opportunity exist for innovation and development in the field of
photovoltaics. As investment in R&D continues, the technology and the case for using
PV for water splitting will continue to strengthen. Another advantage is of the PV
technology is that it supplies DC electricity which is an ideal coupling with the operation
need of an electrolyser.
3.1.2 Plant Operability
The intermittent nature of solar energy introduces complications to powering
electrolysis with PV-electricity. In the field of photovoltaics a capacity utilisation factor
(CUF) is defined as the ratio between the annual energy delivered to the energy that
would be delivered annually under ideal conditions (Equation 3.1)
In summary, the techno-economic assessment of three methods of producing
hydrogen through splitting water has been carried out. For PV-electrolysis plants, the cost
of electricity feedstock was found to have the most influence on the overall cost of
producing hydrogen. Improving the efficiency of electrolysers can reduce this burden
somewhat but developing PV technology play a key role to provide a cheap electricity
feedstock. Photocatalysis and PEC cells are still technically immature but indicate
promise for the future. Current challenges involve discovering new materials that can
prevent electron-hole recombination and utilise a higher proportion of the solar spectrum
to increase STH efficiency which dominates the overall costs in the two technologies.
Once this is achieved it is envisaged that capital costs will become a significant, but not
prohibitive, expense to reduce. Hydrogen production from photocatalysis provided the
most positive economic case with both Type 1 and Type 2 reactors producing hydrogen
at 1.6 $/kgH2 and 3.20 $/kgH2, respectively. Only the Type 4 reactor that utilises PEC
technology was deemed sufficiently competitive, producing hydrogen at 4.05 $/kgH2.
Although more expensive, the inherent separation of gases and greater certainty of
architecture means PEC systems provide a more well-rounded and holistic solution. In
addition, sufficient advancements in the STH efficiency of PEC devices can potentially
reduce the cost to 2.7 $/kgH2. This would bring the economic performance within the US
DOE target and in line with the photocatalytic technology. For these reasons, although
both remain interesting prospects for the future, PEC system are favored. The further
analysis on the different components in a PEC cell underlines the group III-V compounds
have a strong potential to meet the target of high STH efficiency, together with an
appropriate cocatalyst which also works as a protection layer from photocorrosion.
However other photoelectrodes (including metal oxides, sulphides and so on) are not
ruled out at this stage. In total, a highly efficient PEC device with an affordable cost and
stability is the most important in the solar driven water splitting field.

You might also like