You are on page 1of 40

3.

1 Hydrogen Production
Canan Acar, Bahcesehir University, Istanbul, Turkey
Ibrahim Dincer, University of Ontario Institute of Technology, Oshawa, ON, Canada
r 2018 Elsevier Inc. All rights reserved.

3.1.1 Introduction 3
3.1.2 Background 5
3.1.2.1 Hydrogen Production From Fossil Fuels 6
3.1.2.1.1 Production from natural gas 6
3.1.2.1.2 Production from coal 6
3.1.2.1.3 Capture and storage of CO2 7
3.1.2.2 Hydrogen Production From Water Splitting 7
3.1.2.2.1 Water electrolysis 7
3.1.2.2.2 Photoelectrolysis (photolysis) 7
3.1.2.2.3 Photobiological production (biophotolysis) 8
3.1.2.2.4 High-temperature decomposition 8
3.1.2.2.5 Thermochemical water splitting 8
3.1.2.3 Biomass to Hydrogen 8
3.1.2.4 Centralized Hydrogen Production 9
3.1.3 Classification of Hydrogen Production Methods 9
3.1.3.1 Electrolysis 10
3.1.3.2 Plasma Arc Decomposition 12
3.1.3.3 Water Thermolysis 12
3.1.3.4 Thermochemical Water Splitting 12
3.1.3.5 Thermochemical Conversion of Biomass, Gasification, and Biofuel Reforming 13
3.1.3.6 Photovoltaic Electrolysis, Photocatalysis, and Photoelectrochemical Method 13
3.1.3.7 Dark Fermentation 14
3.1.3.8 High Temperature Electrolysis 15
3.1.3.9 Hybrid Thermochemical Cycles 15
3.1.3.10 Coal Gasification 15
3.1.3.11 Fossil Fuel Reforming 16
3.1.3.12 Biophotolysis and Photofermentation 16
3.1.3.13 Artificial Photosynthesis 16
3.1.3.14 Photoelectrolysis 17
3.1.3.15 Summary 17
3.1.4 Efficiency Evaluation of Hydrogen Production Methods 17
3.1.5 Comparative Evaluation of Hydrogen Production Methods 20
3.1.5.1 Comparison of Environmental Impact 20
3.1.5.2 Comparison of Social Cost of Carbon 21
3.1.5.3 Comparison of Hydrogen Production Cost 21
3.1.5.4 Comparison of Energy and Exergy Efficiencies 22
3.1.5.5 Overall Comparison 23
3.1.6 Comparative Evaluation Results of Hydrogen Production Methods 23
3.1.6.1 Hydrogen Production From Electrical Energy 23
3.1.6.2 Hydrogen Production From Thermal Energy 24
3.1.6.3 Hydrogen Production From Photonic Energy 24
3.1.6.4 Hydrogen Production From Hybrid Sources 26
3.1.6.5 Overall Comparison 26
3.1.7 Case Study: Hydrogen Production via Electrolysis 27
3.1.7.1 Alkaline Electrolyzers 28
3.1.7.2 Polymer Electrolyte Membrane Electrolyzers 28
3.1.7.3 Solid Oxide Electrolyzers 29
3.1.7.4 Environmental Impact Analysis of Electrolysis 29
3.1.7.5 Global Warming Potential 30
3.1.7.6 Acidification Potential 32
3.1.8 System Integration in Hydrogen Production 32
3.1.9 Hydrogen Production Challenges 33

Comprehensive Energy Systems, Volume 3 doi:10.1016/B978-0-12-809597-3.00304-7 1


2 Hydrogen Production

3.1.9.1 How Much Hydrogen Is Needed to Meet Our Energy Demand? 34


3.1.9.1.1 Small scale distributed hydrogen production options 34
3.1.9.1.2 Large-scale centralized hydrogen production options 34
3.1.9.1.3 Integrated hydrogen production approach 34
3.1.10 Innovative Hydrogen Production Research at Clean Energy Research Laboratory 35
3.1.10.1 Electricity 35
3.1.10.2 Thermochemical 35
3.1.10.3 Solar 35
3.1.11 Future Directions 36
3.1.12 Closing Remarks 37
References 37
Futher Reading 39
Relevant Websites 39

Nomenclature I_ Solar irradiance, W/m2


_
En Energy flow rate, kW m _ Mass flow rate, kg/s
_
Ex Exergy flow rate, kW n_ Molar flow rate, mol/s-m2
ex Specific exergy, kJ/kg v Frequency, 1/s
G Gibbs free energy, kJ/mol V Voltage
h Planck’s constant, B6.63  10–34 m2-kg/s W _ Electrical power, W
H Enthalpy

Greek Symbols l Wavelength, nm


D Change c Exergy efficiency
Z Energy efficiency

Subscript and Superscripts in Input


aq Solution (aqueous) phase l Liquid phase
ch Chemical out Output
f Free or formation s Solid phase
g Gas phase 1 Standard state
H2 Hydrogen

Acronyms IPCE Incident photon-to-current conversion


AC Alternating current efficiency
AP Acidification potential, g SO2 eq./kg hydrogen LCA Life cycle assessment
produced LCIA Life cycle inventory assessment
APCE Absorbed photon-to-current conversion LHV Lower heating value
efficiency MEA Membrane electrode assembly
ATR Autothermal reforming MTOE Million tons of oil equivalent
CCGT Combined cycle gas turbine NGSR Natural gas steam reforming
CCS Carbon capture and storage or sequestration OECD Organization for Economic Cooperation and
COD Chemical oxygen demand Development
DC Direct current OD Optical density
EIF Environmental impact factor PEC Photoelectrochemical cells
GF Greenization factor PEM Polymer electrode membrane
GHG Greenhouse gases POX Partial oxidation
GWP Global warming potential, g CO2 eq./kg PV Photovoltaic
hydrogen produced SCC Social cost of carbon, $/kg hydrogen produced
HCF Hydrogen content factor SMR Steam methane reforming
HHV Higher heating value SOE Solid oxide electrolyzer
HTE High-temperature electrolysis SPE Solid polymer electrolyte
IAF Integrated assessment framework STH Solar-to-hydrogen
IEA International energy agency TPES Total primary energy supply
Hydrogen Production 3

3.1.1 Introduction

One of the main issues of the 21st century is maintaining to supply the ever growing global energy demand in a sustainable
fashion. There are two reasons behind the significant increase in global energy demand: (1) increase in worldwide population and
(2) rising standards of living. For example, in 2014, 15 TW energy was exhausted by about seven billion people worldwide. These
numbers are predicted to accelerate to 30 TW and nine billion people, respectively, in 2050 [1]. Fig. 1 demonstrates world’s fuel
shares of total primary energy supply (TPES), electricity generation, and the resulting CO2 emissions.
From Fig. 1, it can be understood that 82% of the global TPES came from fossil fuels in 2014. On the other hand, due to
their restricted and nonhomogeneous global reserve supply and distribution, fossil fuels are not estimated to keep sustaining the
ever increasing global energy demand. What is more, fossil fuel resources keep becoming less reachable, while the easily reached
reserves are exhausted. As a result, fossil fuel prices keep on rising because of the loss of easier access to existing reserves and
governmental ambiguities with the regions that have the largest global fossil fuel reserve shares. In conjunction with the
financial problems, greenhouse gas (GHG) (mostly CO2) emissions as a consequence of fossil fuel combustion, and the
contribution to climate change, have been causing momentous environmental concerns. For that reason, substituting traditional
fuels with non-fossil fuel-based energy sources could significantly decrease the GHG emissions and their unfavorable impact on
climate change.
Sustainable energy sources, carriers, and systems could potentially address all the issues mentioned above by lowering the
reliance on fossil fuel reserves and reducing environmentally damaging emissions. Hydrogen is a sustainable energy carrier and can
be used as a fuel in sustainable energy systems due to its zero or negligible end use emissions and the endless possibilities of
producing hydrogen from persistently renewed energy and material sources. Several benefits of hydrogen as an energy carrier and
fuel are: (1) enhanced system efficiencies; (2) production from water with zero emissions; (3) availability; (4) availability of
different storage forms (such as gaseous, liquid, or in metal hydride forms); (5) compatible with extended delivery distances; (6)
easy conversion to many different energy forms; (7) higher HHV and LHV compared to almost all fossil fuels (Table 1). Con-
trariwise, in general, existing hydrogen production options are not commercially mature and current hydrogen production
efficiencies are low, which causes hydrogen production to be expensive [2].
Here, we go further to compare hydrogen with other conventional fuels in terms of environmental impact factor (EIF),
greenization factor (GF) and hydrogen content factor (HCF) to emphasize the importance of hydrogen as a unique option,

Nuclear Hydro Other Other


Oil
Biofuels/waste 5% 2% 1% 4%
5%
10%
Oil Hydro
32% 16%
Nuclear Coal/petroleum
Natural gas 12% 41%
21%

Coal/petroleum Natural gas


29% 22%

(A) (B)

Other
1%

Natural gas
20% Oil
35%

Coal/petroleum
44%

(C)

Fig. 1 Global energy source allocation of (A) total primary energy supply (TPES), (B) electricity production, and (C) CO2 emissions in 2014
(“other” involves geothermal, solar, waste heat, wind, etc.). Data from International Energy Agency. 2016 Key world energy statistics: technical
report. Available from: https://www.iea.org/publications/freepublications/publication/KeyWorld2016.pdf; 2016 [accessed 10.02.16].
4 Hydrogen Production

Table 1 Higher heating value (HHV) and lower heating value (LHV) of
hydrogen and common fossil fuels at 251C and 1 atm

Fuel HHV (kJ/g) LHV (kJ/g)

Hydrogen 141.9 119.9


Methane 55.5 50.0
Gasoline 47.5 44.5
Diesel 44.8 42.5
Methanol 20.0 18.1

Source: Reproduced from Dincer I. Green methods for hydrogen production. Int J Hydrog Energy
2012;37(2):1954–71.

1 4.0

HCF 3.5
0.8
GF 3.0
HCF or GF

EIF 2.5
0.6

EIF
2.0
0.4 1.5

1.0
0.2
0.5

0 0.0
Coal Oil Natural Gas Hydrogen
Fig. 2 Comparison of hydrogen content factor (HCF), greenization factor (GF), and environmental impact factor (EIF) of selected fuels.

through the following equations:


kg CO2 product of combustion reaction
EIF ¼ ð1Þ
kg fuel

EIFmax  EIF
GF ¼ ð2Þ
EIFmax

kg of H2 in the fuel
HCF ¼ ð3Þ
kg fuel

where EIFmax is the maximum value of EIF among the evaluated options. In this specific case with 3.6, coal is selected as the EIFmax.
Fig. 2 shows that as the hydrogen content (HCF) increases, the fuels turn out to be greener (i.e., higher GF) and their negative
environmental impact (EIF) declines. This shows a strong benefit of hydrogen in regards to minimizing carbon-based emissions.
With the intention of taking maximum amount of benefits from the hydrogen energy systems, its production must be from
renewable, clean, abundant, reliable, affordable, and available energy and material resources. There are quite a lot of investigations
in the literature (such as Refs. [3–5]) concentrating on how hydrogen has a potential to become one of the most beneficial
solutions to key energy issues by supporting the global sustainability. Among the alternative hydrogen production methods
investigated in the literature, natural gas steam reforming (NGSR) is the most generally preferred one due to its already mature
technology and heavy use in the industry. However, steam reforming causes harmful GHG emissions. Currently, about half of the
worldwide hydrogen supply comes from NGSR. The rest is distributed as the following: 30% from oil reforming, 18% from coal
gasification, 3.9% from water electrolysis, and 0.1% from other sources [6]. With the purpose of removing the negative impacts of
fossil fuel use on the environment, economy, and the health of living beings, hydrogen has to be produced from affordable, clean,
and abundant resources and systems [7,8]; this is considered as the greenization of hydrogen production.
Green hydrogen production options are not presently very efficient or affordable. For example, existing information in the
literature shows that PV electrolysis is currently expensive with production costs higher than $5/kg H2. In addition, most novel solar-
based hydrogen production systems are at low efficiencies; some of them have about 5% energy and exergy efficiencies with current
state of technologies. However, the literature suggests that it is only a matter of time for novel hydrogen production methods to
evolve as green, highly efficient, and a lot cheaper with the introduction of advanced materials innovative technologies [9].
Hydrogen Production 5

Hydrogen production from renewable and abundant energy and material resources has formerly been investigated in the
literature by numerous researchers. High- and low-temperature water electrolysis, thermochemical water dissociation, and pho-
tolysis have been analyzed by Lodhi [10]. This study has been regarded as one of the early investigations of novel
hydrogen production systems. Following this study, Lodhi [11] has categorized hydro, nuclear, ocean (or sea), solar, and wind
energy as principal green energy resources for hydrogen production. Lodhi [11] has listed green material resources for hydrogen
production as water (fresh or sea/ocean water), hydrogen sulfide, and biomass. Hydrogen production methods can be regarded as
“green” depending on their primary energy and material resources [12]. Lemus and Duart [13] have investigated cost of centralized
and distributed hydrogen production along with some hydrogen distribution challenges such as compression, storage, and
delivery.
Another novel production method for hydrogen is to mimic the photosynthesis process. Alternative hydrogen production
methods developed by mimicking photosynthesis have been listed by Alstrum-Acevedo et al. [14]. Tanksale et al. [15] have
reviewed biomass-based catalytic hydrogen production methods, such as gasification and pyrolysis. Acar and Dincer [2] have
comparatively investigated the cost, emissions, and efficiencies of steam methane reforming (SMR), coal gasification, solar and
wind-based electrolysis, biomass gasification, Cu–Cl and S–I based thermochemical water splitting cycles, and high-temperature
electrolysis (HTE).
In this chapter, a wide ranging review of hydrogen production alternatives from renewable and nonrenewable energy and
material resources is presented. Here, the selected hydrogen production alternatives are discussed in detail and they are com-
paratively investigated. Electrical, thermal, biochemical, photonic, electrothermal, photoelectric, and photobiochemical are the
principal energy resources evaluated in this chapter. As a case study, 20 hydrogen production technologies are evaluated
regarding their energy and exergy efficiencies, hydrogen production costs, emissions, and SCC. This chapter also outlines the
characteristics of traditional and innovative hydrogen production systems, clarifies that the complete evolution to clean hydrogen
production might take some time, and lists the possible advantages of such a successful switch to a fully developed hydrogen
economy.

3.1.2 Background

Extensive global utilization of hydrogen as an energy carrier and fuel can potentially solve issues related to energy supply security,
climate change, and resulting health concerns. Hydrogen can be extracted by using a wide range of available, abundant, and
reliable primary energy and material resources. These resources include fossil fuels, renewables, biomass, and nuclear.
In spite of these very promising advantages, there are many challenges to tackle before taking full advantage of the hydrogen
economy. Different than fossil fuels, hydrogen does not have a current large-scale production, delivery, storage, and end use
infrastructure. Therefore the first requirement is to build a solid, reliable, convenient, efficient, and clean infrastructure that
requires major investments. Even though current production, storage, and delivery methods for hydrogen are presently used by the
chemical and refining industries, present hydrogen production, storage, and conversion technologies are significantly expensive for
extensive energy related end use requirements. In addition, prevailing standards and policies do not highlight the external cost
reduction related to environmental and energy security, which would be the economic advantage of widespread hydrogen use as
an energy carrier and fuel. Table 2 presents the key issues that support and inhibit the wide use of hydrogen in energy systems.
Making hydrogen as the future energy carrier and fuel requires an extraordinary amount of sustained and coordinated activities
by diverse stakeholders including academia, industry, governments, and the general public. Knowing the need to develop
innovative, clean, reliable, affordable, and efficient hydrogen production systems, all stakeholders from the local and regional
governments, industries, academia, research institutions, and the public should initiate an effective vision and roadmap for the
production of hydrogen and implementation of hydrogen energy systems to integrate the ideas and perspectives of an extensive
range of stakeholders.

Table 2 Summary of key drivers to support and inhibit when developing innovative hydrogen production systems

Support Inhibit Both support and inhibit

Energy security and fewer fuel Difficulties in building and sustaining Innovations in hydrogen and alternative
imports consensus on energy policies technologies
Greenhouse gas (GHG) emissions Lack of an existing hydrogen infrastructure Lower cost fossil fuels and their rapid
and global climate change Lack of commercial and affordable hydrogen depletion
Increasing global population and production,a storage, and end use Consumer desire for clean, affordable,
living standards technologies convenient, reliable, abundant, and available
Need for novel, clean, affordable, Hydrogen safety issues energy sources
reliable, efficient energy systems Cost of innovative hydrogen energy systems
Air quality and its impacts on the
health of living beings
a
Despite the fact that large amounts of hydrogen can be produced relatively cheaply by steam methane reforming, this method requires limited fossil fuels (i.e., natural gas) as energy
source and also emits greenhouse gas (mainly CO2).
6 Hydrogen Production

The aim of developing sustainable ways to produce hydrogen is to find a shared idea for the current and future hydrogen energy
systems. It is also aimed to develop a mutual time frame within which this hydrogen economy vision could be accomplished. And it is
also important to identify the major milestones for accomplishing a successful transition to hydrogen energy systems. Hydrogen energy
systems’ vision could be summarized by the subsequent declaration: hydrogen is a globally promising future clean energy option. It is
adaptable, reasonably priced, safe, locally available, and can be utilized in almost every aspect of life and in all areas of the world. The
most important driving forces and motivation for developing innovative hydrogen production systems can be summarized as:

• Hydrogen, as an energy carrier and a fuel, occupies an important part in our global energy supply mix; it could decrease the
dependence on fuel imports and pollutant emissions.
• Evolution to hydrogen energy systems has already started and we are getting closer to fully developed hydrogen energy systems
integrated in our daily lives.
• Hydrogen energy systems, together with all components, need innovative approaches.
• There are “chicken and egg” problems with introducing hydrogen to existing markets.
• Local, regional, and state level governments should develop and maintain reliable and steady energy policies in order to make
hydrogen as a preferred energy option.
• Strong collaborations from the industry, academia, governments, and the public are needed to find innovative methods to
produce and utilize hydrogen energy.
• A global hydrogen energy vision is required to strengthen research, development, demonstration, outreach, and policies
regarding hydrogen production, delivery, storage, and end use.

3.1.2.1 Hydrogen Production From Fossil Fuels


Most of the fossil fuel sources can be used to produce hydrogen in many different ways with varying complexities. In this chapter,
hydrogen production from natural gas and coal are presented. When fossil fuels are used to produce hydrogen, carbon dioxide is
released as a byproduct and it should be captured to guarantee clean hydrogen production. Fossil fuel-based hydrogen production
methods’ feasibilities depend on the scale of operation and whether a plant has centralized or distributed operation.

3.1.2.1.1 Production from natural gas


Currently available natural gas-based hydrogen production methods can be listed as:

• SMR
• partial oxidation (POX)
• autothermal reforming (ATR)
Even though there are several novel natural gas-based hydrogen production methods available, none of them is commercia-
lized yet. SMR includes an endothermic reaction between methane and steam, which yields hydrogen and carbon monoxide (Eq.
(4)). The heat source of this endothermic reaction is the methane gas input to the reaction chamber. This reaction generally takes
place at temperatures between 700 and 8501C and within a pressure range of 3 to 25 bar. As a result of this reaction, output gas
includes around 12% CO by volume, which is then used to react with steam and converted to CO2 and H2 via the water gas shift
reaction (Eq. (5)).
CH4 þ H2 O þ heat-CO þ 3H2 ð4Þ

CO þ H2 O-CO2 þ H2 þ heat ð5Þ


In POX, natural gas partially combusted with pure oxygen to produce hydrogen to produce carbon monoxide and hydrogen
(Eq. (6)). This is an exothermic reaction so the reactor does not require a heat source. The resulting CO is then converted to H2 as
shown in Eq. (5).
1
CH4 þ O2 -CO þ 2H2 þ heat ð6Þ
2
ATR is a combination of steam reforming (Eq. (4)) and POX (Eq. (6)). The overall reaction is exothermic and its output
temperature ranges between 950 and 11001C. The reactor pressure can go up to 100 bar. The product CO is converted to H2
through the water gas shift reaction (Eq. (5)). The output gas purification considerably increases the operation costs and also
decreases the overall system efficiencies. Each one of the above processes has certain benefits and challenges, which are sum-
marized in Table 3.

3.1.2.1.2 Production from coal


Hydrogen can be produced from coal through a variety of gasification processes (e.g., fixed bed, fluidized bed, or entrained flow).
In practice, high-temperature entrained flow processes are favored to maximize carbon conversion to gas, thus avoiding the
formation of significant amounts of char, tars, and phenols. A typical reaction for the process is given in Eq. (7) in which carbon is
converted to carbon monoxide and hydrogen.
CðsÞ þ H2 O þ heat-CO þ H2 ð7Þ
Hydrogen Production 7

Table 3 Technology comparison of natural gas-based hydrogen production

Technology Benefits Challenges

Steam methane reforming (SMR) High efficiency Complex system


Emissions Sensitive to natural gas qualities
Costs for large units
Autothermal reforming (ATR) or partial Smaller size Lower efficiency
oxidation (POX) Costs for small units H2 purification
Simple system Emissions and flaring

Since this reaction is endothermic, additional heat is required, as with methane reforming. The CO is further converted to CO2
and H2 through the water gas shift reaction, described in Eq. (5). Coal-based hydrogen production is a commercially developed
process. However, it is more difficult compared to natural gas-based hydrogen production. As a result, coal-based hydrogen pro-
duction is more expensive. On the other hand, coal is abundant in most regions of the world and it is expected to be utilized as an
energy source anyway. Therefore it is meaningful to search the innovative and clean methods to utilize coal for hydrogen production.

3.1.2.1.3 Capture and storage of CO2


CO2 is the main emission of all fossil fuel-based hydrogen production processes. The amount of CO2 emissions depend on the
hydrogen content of the fossil fuel source. In order to develop a hydrogen production system with zero or minimum emissions,
the CO2 emissions must be captured and stored (also known as decarbonization). There are three ways to capture and store CO2 in
a fossil fuel-based hydrogen production system:

• Postcombustion: CO2 emissions can be eliminated from the combustion chamber’s exhaust. This process is mostly used in
conventional steam turbines or combined cycle gas turbine (CCGT) power plants, and is generally conducted by the “amine”
process. Here, the emissions stream includes considerable quantities of nitrogen and traces of nitrogen oxides along with steam,
CO2, and CO.
• Precombustion: CO2 is captured as hydrogen being produced via any of the reactions explained earlier.
• Oxyfuel combustion: in this process, with pure oxygen is used to react with fossil fuels (instead of air). As a result, almost only
CO2 and steam are the emissions, and CO2 can be simply captured by condensing the steam in the mixture.

3.1.2.2 Hydrogen Production From Water Splitting


Hydrogen can be produced via dissociating water molecules via several different processes. This section summarizes high- and low-
temperature electrolysis, photoelectrolysis, and photobiological hydrogen production.

3.1.2.2.1 Water electrolysis


Water electrolysis is the process whereby water is split into hydrogen and oxygen through the application of electrical energy, as in
Eq. (8). The total energy that is needed for water electrolysis is increasing slightly with temperature, while the required electrical energy
decreases. A HTE process might, therefore, be preferable when high-temperature heat is available as waste heat from other processes.
This is especially important globally, as most of the electricity produced is based on fossil energy sources with relatively low efficiencies.

1
H2 O þ electricity-H2 þ O2 ð8Þ
2

3.1.2.2.2 Photoelectrolysis (photolysis)


PV powered electrolyzers are already available in the market. These options provide a certain degree of flexibility since the product
can either be electricity or hydrogen. Direct photoelectrolysis is a promising alternative to PV-based electrolysis systems. This is
because of the fact that in photoelectrolysis solar-to-hydrogen (STH) production occurs directly in a single apparatus without the
need of another device for electricity generation. Photoelectrolysis can reduce hydrogen production costs and become cheaper than
the two step PV-based electrolysis.
Significant research and development activities are required to fully take advantage of advanced materials science and develop
innovative systems engineering approaches. Photoelectrochemical (PEC) hydrogen production is an example of these innovative
approaches. PECs are presently being investigated worldwide. At least 13 Organization for Economic Cooperation and Devel-
opment (OECD) countries carry on research and development on PEC systems. There are four main PEC systems that are currently
being investigated in the literature: (1) two photon tandem cells, (2) monolithic multijunction cells, (3) dual bed redox cells, and
(4) one pot two step cells. Systems (1) and (2) use thin film coated glass submerged in water. Systems (3) and (4) use photoactive
powder catalysts dissolved in water. There are several small scale PEC demonstrations presented in the literature; so far these
systems could reach up to 16% STH conversion efficiencies.
The key challenges to advance PEC cell innovation toward the market concern progress in materials science and engineering. It
is very important to develop photoelectrode materials and their processing technologies with high efficiency (performance) and
8 Hydrogen Production

corrosion resistance (longevity) characteristics, paving the path toward smart system integration and engineering. Since no “ideal”
photoelectrode material commercially exists for water splitting, tailored materials have to be engineered.
Combinatorial chemistry approaches offer fast-tracking experimental options for the necessary materials screening, while
modeling capabilities of photooxidation based on quantum transition theory need to be developed. Most important, there is a
need for fundamental research on semiconductor doping for band gap shifting and surface chemistry modification, including
studies of the associated effects on both surface and bulk semiconducting properties. Corrosion and photocorrosion resistance
present further significant research and development challenges to be addressed, with most of the promising materials options at
hand. Current-matching between anode and cathode, in addition to ohmic resistance minimization, requires considerable systems
design as well as sophisticated engineering solutions. Optimization of fluid dynamics (with its effects on mass and energy transfer)
and gas collection and handling (with its effects on operational safety) will demand major conceptual and application specific
research and development attention.

3.1.2.2.3 Photobiological production (biophotolysis)


Photobiological production of hydrogen is based on two steps: photosynthesis (Eq. (9)) and hydrogen production catalyzed by
hydrogenases (Eq. (10)) in, for example, green algae and cyanobacteria. Long-term basic and applied research is needed in this
area, but if successful, a long-term solution for renewable hydrogen production will result. It is of vital importance to understand
the natural processes and the genetic regulations of H2 production. Metabolic and genetic engineering may be used to demonstrate
the process in larger bioreactors. Another option is to reproduce the two steps using artificial photosynthesis.
Photosynthesis : 2H2 O-4Hþ þ 4e þ O2 ð9Þ

Hydrogen Production : 4Hþ þ 4e -2H2 ð10Þ

3.1.2.2.4 High-temperature decomposition


High-temperature splitting of water occurs at about 30001C. At this temperature, 10% of the water is decomposed and the
remaining 90% can be recycled. To reduce the temperature, other processes for high-temperature splitting of water have been
suggested:

• Thermochemical cycles.
• Hybrid systems coupling thermal decomposition and electrolytic decomposition.
• Direct catalytic decomposition of water with separation via a ceramic membrane (thermophysic cycle).
• Plasmachemical decomposition of water in a double-stage CO2 cycle.

For these processes, efficiencies above 50% can be expected and could possibly lead to a major decrease of hydrogen pro-
duction costs. The main technical issues for these high-temperature processes relate to materials development for corrosion
resistance at high temperatures, high-temperature membrane and separation processes, heat exchangers, and heat storage media.
Design aspects and safety are also important for high-temperature processes.

3.1.2.2.5 Thermochemical water splitting


Thermochemical water splitting is the conversion of water into hydrogen and oxygen by a series of thermally driven chemical
reactions. Thermochemical water splitting cycles have been known for the past 35 years. They were extensively studied in the late
1970s and 1980s, but have been of little interest in the past 10 years. While there is no question about the technical feasibility and
the potential for high efficiency, cycles with proven low cost and high efficiency have yet to be developed commercially. An
example of a thermochemical process is the iodine/sulfur cycle, outlined in Eqs. (11)–(14). For this process, the research and
development needs are to capture the thermally split H2, to avoid side reactions and to eliminate the use of noxious substances.
The corrosion problems associated with the handling of such materials are likely to be extremely serious.
1
At 8501C : H2 SO4 -SO2 þ H2 O þ O2 ð11Þ
2

At 120 1C : I2 þ SO2 þ 2H2 O-H2 SO4 þ 2HI ð12Þ

At 4501C : 2HI-I2 þ H2 ð13Þ

1
Overall : H2 O-H2 þ O2 ð14Þ
2

3.1.2.3 Biomass to Hydrogen


In biomass conversion processes, a hydrogen containing gas is normally produced in a manner similar to the gasification of coal,
as in Eq. (7). However, no commercial plants exist to produce hydrogen from biomass. Currently, the pathways followed are
steam gasification (direct or indirect), entrained flow gasification, and more advanced concepts, such as gasification in supercritical
Hydrogen Production 9

water, application of thermochemical cycles, or the conversion of intermediates (e.g., ethanol, bio-oil, or torrified wood). None of
the concepts have reached a demonstration phase for hydrogen production.
Biomass gasification is in the research and development area shared between H2 production and biofuels production. Gasi-
fication and pyrolysis are considered the most promising medium term technologies for the commercialization of H2 production
from biomass. In terms of its energy requirements, the drying of biomass might not be justifiable; therefore, other pathways based
on wet biomass are being sought as well.
Biomass feedstocks are unrefined products with inconsistent quality and poor quality control. The production methods vary
according to crop type, location, and climatic variations. Erratic fuels have contributed to the difficulties in technological inno-
vation, since less homogenous and low quality fuels need more sophisticated conversion systems. There is a need to rationalize the
production and preparation of fuel to produce more consistent, higher quality fuels that can be described by common standards.
Large-scale systems tend to be suitable for cheaper and lower quality fuels, while smaller plants tend to require higher levels of fuel
quality and better fuel homogeneity. A better understanding of this relationship, and the specific tolerances that each technology
can accommodate, is needed.
Several developments are needed to improve the economics of production processes and the logistics of handling a biomass
feedstock:

• Feed preparation and identifying the characteristics of feedstocks that will allow the technologies to develop.
• Gasification of biomass: this is not specific for hydrogen, but relates to general biomass and renewables pathways and research.
• Raw gas handling and clean-up.
• Interface issues and system integration. One should also investigate the relationship between the production scale and the fuel
quality requirements and tolerances that can be accommodated for the respective technologies.

3.1.2.4 Centralized Hydrogen Production


Large scale, industrial hydrogen production from all fossil energy sources can be considered a commercial technology for industrial
purposes, though not yet for utilities. Hydrogen production at a large scale has the potential for relatively low unit costs, although
the hydrogen production cost from natural gas in medium-sized plants may be reduced toward the cost of large-scale production.
An important challenge is to decarbonize the hydrogen production process. CO2 capture and storage options are not fully
technically and commercially proven. They require research and development on absorption or separation processes and process
line-up, as well as acceptance for CO2 storage. It is also important to increase plant efficiency, reduce capital costs and enhance
reliability and operating flexibility.
Further research and development activities are particularly needed on hydrogen purification (to produce H2 suitable for fuel
cells) and on gas separation (to separate hydrogen or CO2 from gas mixtures). This involves the development of catalysts,
adsorption materials, and gas separation membranes for the production and purification of hydrogen. Hydrogen and power can
be coproduced in integrated gasification combined cycle (IGCC) plants. The IGCC plant is the most advanced and efficient
solution in which the carbon in the fuel is removed and the hydrogen is produced in a precombustion process.
However, successful centralized hydrogen production requires large market demand, as well as the construction of a new
hydrogen transmission and distribution infrastructure and pipeline for CO2 storage. In the future, centralized hydrogen pro-
duction from high-temperature processes based on renewable energy and waste heat can also be an option to enhance sustain-
ability and remove the need for capture and storage of CO2.

3.1.3 Classification of Hydrogen Production Methods

In order to identify pathways for hydrogen production one needs to inventory the natural resources of hydrogen, the available
sources of energy that can be used to extract hydrogen from natural resources, and the applicable methods of hydrogen pro-
duction. These items are summarized in Fig. 3. The hydrogen-containing natural resources are water, fossil fuels, biomass,
hydrogen sulfide, and anthropogenic wastes as indicated in Fig. 3. Municipal sewage waters containing urea, farming wastes like
manure, crops residues, etc. (which are sources of biogas), other wastes that generate landfill gas, recycled plastic and cellulosic
materials, etc., are all anthropogenic wastes from which hydrogen can be extracted.
Sustainable energy is required to extract hydrogen from any resource in a clean, nonpolluting manner. Fig. 3 lists the main
energy sources that can be considered sustainable: solar, hydro, ocean thermal, tidal (including ocean currents), wind, biomass,
geothermal, and nuclear. The careful use of any of these sources can generate electricity and/or high-temperature heat and/or
nuclear radiation without or with minor environmental impact. Such energy is used for hydrogen production via one of the
methods, as listed in Fig. 3. These methods are categorized in six classes, namely, electrochemical, thermochemical, photo-
chemical, radiochemical, biochemical, and hybrid (where hybrid methods refer to integrated systems that use any kind of
combination of the first five listed hydrogen production methods, e.g., electrophotochemical, photobiochemical,
electrothermochemical, etc.).
Based on the above analysis, there are five possible pathways to generate hydrogen in a sustainable manner. As indicated in
Fig. 3, these are: water splitting, fossil hydrocarbons decarbonization, and hydrogen sulfide decomposition, biomass conversion to
hydrogen, and extraction of hydrogen from waste materials resulted from the anthropogenic activity. Each pathway corresponds to
10 Hydrogen Production

Water Solar

Fossil fuels Hydro


Hydrogen containing
material resources Biomass Ocean thermal

Hydrogen sulfide Tidal


Sustainable energy
Anthropogenic wastes sources Wind

Biomass

Geothermal

Nuclear

Electrochemical Fossil fuel decarbonization

Thermochemical Hydrogen sulfide decomposition

Photochemical Biomass conversion


Hydrogen production Sustainable hydrogen
methods production pathways
Radiochemical Hydrogen extraction from waste

Biochemical Water splitting

Hybrid Hybrid

Fig. 3 Classification of hydrogen production based on material resources, sustainable energy sources, existing methods, and sustainable
pathways.

a natural resource (including the anthropogenic waste) from which hydrogen can be extracted. For any of the pathways the use of a
specific combination of sustainable energy (see Fig. 3) and hydrogen production method (see Fig. 3) is possible. In the subsequent
sections of this chapter, the main options are addressed.
Hydrogen is abundant in nature and exists in many compounds such as water, biomass, hydrogen sulfide, and fossil fuels.
Table 4 shows an outline of 19 hydrogen production methods with their principal energy and material resources.
Among the energy sources mentioned in Table 4, electrical and thermal energies can be derived either from fossil fuels with
carbon capture and storage (CCS), or renewable resources (e.g., hydro, ocean, solar, thermal, wave, wind, etc.), or biomass, or
nuclear, or energy recovered from different processes. The source of photonic energy is solar radiation and biochemical energy is
obtained from organic substances, such as biomass. Other than the four key principal energy resources presented in Table 4 (i.e.,
biochemical, electrical, photonic, and thermal), there are some combinations of these principal energy resources including
electrical-photonic (photoelectrical), electrical-thermal (electrothermal), photonic-biochemical (photobiochemical), etc. Water,
biomass, and fossil fuels are the material resources investigated in this section.

3.1.3.1 Electrolysis
At present, the most fundamental industrial water splitting process for nearly pure hydrogen generation is water electrolysis. There
are several types of electrolyzers and they are expected to become more important as new technologies and materials evolve. Water
electrolysis relies on the transfer of electrons as a result of the applied external circuit. Alkaline, polymer membrane (PEM), and
solid oxide electrolyzers (SOE) are the main electrolytic hydrogen generation methods. Table 5 is an overview of the characteristic
properties of alkaline electrolyzers, PEM, and SOE. From the specifications presented in Table 5, voltage efficiency and the current
density are the most significant ones. Efficiency of an electrolysis process is determined by comparing the ideal and real energy
requirements of the electrolytic reactions.
To increase the current density and the rate of electrolysis reactions, catalysts are often preferred in electrolysis processes.
Among the commercially available heterogeneous catalysts, platinum is one of the most frequently preferred ones. Platinum is
generally coated on the surface of either one or both of the electrodes. In addition, some homogeneous catalysts are also selected
to activate or speed up electrolysis reactions. Homogeneous catalysts have high turnover rates. Therefore they are generally cheaper
compared to the heterogeneous catalysts. Some studies in the literature show that some homogeneous catalysts have turnover rates
around 2.4 mol H2/mol catalyst [16].
Electrolyzers, particularly PEM electrolyzers, are very sensitive to the impurities in water. Therefore desalination and demi-
neralization have to be conducted before the electrolyzer. For example, if saline water (such as sea water) is used in an electrolyzer,
the electrolyzer tends to generate chlorine instead of oxygen. In the literature, there are numerous solutions are suggested to block
side reactions (such as the chlorine generation one) in electrolysis. The most common solution in the literature is the use of ion
selective membranes, which could operate well with impure water. El-Bassuoni et al. [17] have first proposed the use of ion
Table 4 Overview of hydrogen production methods by primary energy and material source

Method Primary source Brief description

Energy Material

M1 Electrolysis Electrical Water Direct current (DC) is used to split water into O2 and H2
M2 Plasma arc decomposition Fossil fuels Cleaned natural gas is passed through plasma arc to generate H2 and carbon soot
M3 Thermolysis Thermal Water Thermal decomposition of steam above 2500K
M4 Thermochemical processes Water splitting Water Cyclical reactions (net reaction: water splitting into H2)
M5 Biomass conversion Biomass Thermocatalytic conversion
M6 Gasification Biomass conversion to syngas
M7 Reforming Biofuel conversion to H2
M8 Photovoltaic (PV)-based electrolysis Photonic Water PV panels supports the electrolysis reaction
M9 Photocatalysis Direct water splitting by photocatalyst
M10 Photoelectrochemical (PEC) cells A hybrid cell simultaneously produces current and voltage upon absorption of light
M11 Dark fermentation Biochemical Biomass Biological systems are used to generate H2 in the absence of light
M12 High-temperature electrolysis (HTE) Electrical þ thermal Water Electrical and thermal energy are used together to drive water splitting at high temperatures
M13 Hybrid thermochemical cycles Electrical and thermal energy together in cyclical reactions
M14 Coal gasification Conversion of coal to syngas
M15 Fossil fuel reforming Fossil fuels are converted to H2 and CO2
M16 Biophotolysis Photonic þ biochemical Biomass þ water Biological systems (microbes, bacteria, etc.) produce H2
M17 Photofermentation Fermentation process activated by exposure to light
M18 Artificial photosynthesis Mimicking photosynthesis to produce H2
M19 Photoelectrolysis Electrical þ photonic Water Photoelectrodes and external electricity drive water splitting

Source: Reproduced from Dincer I. Green methods for hydrogen production. Int J Hydrog Energy 2012;37(2):1954–71.

Hydrogen Production
11
12 Hydrogen Production

Table 5 Typical specifications of alkaline, polymer membrane (PEM), and solid oxide electrolyzers (SOE)

Specification Alkaline PEM SOE

Technology maturity State of the art Demonstration R&D


Cell temperature, 1C 60–80 50–80 900–1000
Cell pressure, bar o30 o30 o30
Current density, A/cm2 0.2–0.4 0.6–2.0 0.3–1.0
Cell voltage, V 1.8–2.4 1.8–2.2 0.95–1.3
Power density, W/cm2 Up to 1.0 Up to 4.4 –
Voltage efficiency, % 62–82 67–82 81–86
Specific system energy consumption, kWh/Nm3 4.5–7.0 4.5–7.5 2.5–3.5
Hydrogen production, Nm3/h o760 o30 –
Stack lifetime, h o90,000 o20,000 o40,000
System lifetime, years 20–30 10–20 –
Hydrogen purity, % 499.8 99.999 –
Cold start-up time, min 15 o15 460

Source: Reproduced from Bhandari R, Trudewind CA, Zapp P. Life cycle assessment of hydrogen production via electrolysis – a review. J Clean Prod
2014;85:151–63.

selective membranes in electrolyzers. Ni et al. [18] have shown that magnesium as a catalyst enables the oxygen evolution reaction
in electrolyzers rather than the chlorine generation one.

3.1.3.2 Plasma Arc Decomposition


Plasma is an ionized state of matter that comprises electrons in an excited state along with other atomic substances. Plasma can
potentially be utilized as the preferred medium in processes with high voltage electric current release since it contains electrically
charged particles. In plasma arc decomposition, natural gas (typically methane) is split into hydrogen and carbon black (soot) due
to chemical activity of the thermal plasma. Carbon black is in solid phase, therefore, it stays at the bottom of the reactor as
hydrogen is generated and separately accumulated in gas phase. The plasma arc decomposition reaction of natural gas can be
written as
CH4 -CðsÞ þ 2H2ðgÞ ; DH ¼ 74:6 MJ=kmol ð15Þ

Eq. (15) has been investigated by Fulcheri et al. [19]. In their study, the plasma arc decomposition reaction contains three
electrodes attached to a three phase voltage supplier. Plasma gas is supplied to two of the three electrodes and methane is
introduced from the top inlet of the reactor. Fulcheri et al. [19] have stated that their product is 100% pure hydrogen with no CO2
emissions. It should be noted that in their system, the carbon black stays at the bottom of the reactor in solid phase. Plasma arc
decomposition is also known as high-temperature pyrolysis. Gaudernack and Lynum [20] have stated that plasma arc decom-
position can potentially lower hydrogen production costs by no less than 5% cheaper than large-scale SMR with carbon capture
and sequestration.

3.1.3.3 Water Thermolysis


Water thermolysis is also referred as one-step thermal splitting of water. The water thermolysis reaction is:
heat 1
H2 O ⟶ H2 þ O2 ð16Þ
2
In order to have a reaction that proceeds very close to completion or maximum degree of dissociation, a high-temperature heat
source exceeding 2500K is needed. For example, at a temperature and pressure of 3000K and 1 bar, this reaction (Eq. (16)) has a
degree of dissociation around 65%. Separating the product gases oxygen and hydrogen is one of the critical issues of this method.
There are no semipermeable membranes that are active and resistant to heat at temperatures above 2500K. For that reason, the
product gases need to be cooled down prior to the separation unit. Baykara et al. [21] have conducted the experimental inves-
tigation of a solar powered water thermolysis unit and reached to 90% of the reaction equilibrium at a residence time of 1 ms at a
temperature around 2500K. In addition, the authors’ results have showed that if the resulting oxygen and hydrogen are quickly (in
few milliseconds) cooled down to somewhere around 1500 and 2000K; hydrogen and oxygen recombination can possibly be
eliminated via palladium membranes, which separate hydrogen in an effective manner.

3.1.3.4 Thermochemical Water Splitting


The key advantage of thermochemical water splitting cycles is that they do not require any catalysts to initiate or speed up the
individual reactions in any step during the process. Another benefit is the fact that all chemicals used in the thermochemical water
splitting cycle are recovered and recycled with the exception of water, which is the material resource consumed when producing
Hydrogen Production 13

hydrogen. There are many other advantages of thermochemical water splitting cycles for hydrogen production cycles, such as (1)
product gas separation without the need for special membranes, (2) moderate operating temperatures between 600 and 1200K,
and (3) minimum or very low power input requirement.
Balta et al. [22] have reviewed the literature studies focusing on thermochemical water splitting cycles for hydrogen production.
Among the available thermochemical water splitting cycles, the S–I cycle is one of the options that is regarded as technically
feasible. S–I cycles are being fully built and tested in many countries, such as Japan and the United States. Contrariwise, the
financial feasibility of thermochemical water splitting cycles has to be demonstrated. S–I cycle can be broken down into individual
steps, each one having a different individual reaction. The first reaction is an endothermic dissociation reaction as shown below:
heatð3005001CÞ
H2 SO4ðaqÞ ⟶H2 OðgÞ þ SO3ðgÞ ð17Þ

Steam and SO3, the product gases of the sulfuric acid dissociation reaction, are separately collected and further heated to
temperatures around 800–9001C. Next, SO3 gas is dissociated via another endothermic reaction:
heatð8009001CÞ 1
SO3ðgÞ ⟶ O2ðgÞ þ SO2ðgÞ ð18Þ
2
O2 and SO2, the product gases of the SO3 dissociation reaction, are separately collected. Next, SO2 gas has an exothermic reaction
with iodine and water. This reaction spontaneously starts and proceeds at low temperatures as
SO2ðgÞ þI2ðgÞ þ2H2 OðlÞ -2HIðgÞ þH2 SO4ðaqÞ ð19Þ
here, the product gas H2SO4 is recycled back to the first reaction. And the last step is the thermal decomposition of HI into H2 at
temperatures between 425 and 4501C:
heatð4254501CÞ
2HIðgÞ ⟶H2ðgÞ þ I2ðgÞ ð20Þ

There are no side reactions or undesired byproducts in S–I based thermochemical water splitting cycle. It is somewhat
uncomplicated to take the reaction byproducts in each step and reuse them again in appropriate reactions stated in Eqs. (17)–(20).
Thermochemical water splitting cycles, such as the S–I cycle explained above, have relatively high operating temperature
requirements and most of the reactions are endothermic so the heat requirement of these cycles is very high. Therefore it is a
challenge to support the thermal energy requirements of these cycles in a sustainable manner. Heat recovered from many
industrial processes and novel energy sources such as biomass, solar, and nuclear can support thermochemical water splitting
cycles in a sustainable way.

3.1.3.5 Thermochemical Conversion of Biomass, Gasification, and Biofuel Reforming


Biobased sources (such as biomass and biofuel) can be used to produce hydrogen after proper treatment in order to keep the
energy and material resources’ humidity at a certain level. This is usually done by either various drying processes or supercritical
steam gasification. Biomass (e.g., sugar cane, wood sawdust, etc.) can be converted into hydrogen via the following reaction:
heat
aCl Hm On þ bH2 O ⟶ aH2 þ bCO þ cCO2 þ dCH4 þ eC þ f Tar ð21Þ
here, ClHmOn is the most commonly used chemical formula for various types of biomass. Tar is the unwanted byproduct of the
biomass conversion reaction as it has negative impacts on the overall hydrogen production process (such as contamination and
slugging). In the literature, there are various catalysts with proven ability to regulate, minimalize, and eliminate tar production in
biomass processing reactions (Eq. (21)).
In cases where solid biomass is used for hydrogen production, the subsequent biomass gasification reaction takes place:
high temperature heat
y 
Cx Hy þ xH2 O⟶ þ x H2 þ xCO ð22Þ
2
In biomass gasification processes, fixed, moveable, and fluidized bed reactors are the gasifier types for hydrogen production.
The biomass gasification process can either be thermal or autothermal, depending on the heat input requirements. In autothermal
gasification, the heat requirement is met via a POX reaction in the gasification reactor. When liquid phase biomass (usually
regarded as biofuels such as ethanol, methanol, etc.) is used, the hydrogen production process is very similar to the thermo-
chemical reactions described here in earlier sections.

3.1.3.6 Photovoltaic Electrolysis, Photocatalysis, and Photoelectrochemical Method


Hydrogen production via PV supported electrolysis includes two steps. The first step is the generation of electricity, which is the
conversion of photonic energy to electrical energy via PV panels. And in the next step, electrical energy is converted to hydrogen
energy via electrochemical conversion called electrolysis. In this method, PV panels, converters between alternating current (AC)
and direct current (DC) power, batteries, and hydrogen storage mediums are used in addition to the electrolyzer. PV supported
electrolysis is currently a quite expensive method for producing hydrogen in a sustainable manner. For instance, at the present
technological advancement level, hydrogen produced via PV electrolysis is more than 20 times expensive than hydrogen produced
14 Hydrogen Production

by fossil fuel processing. On the other hand, the cost of PV-based electrolysis has been steadily declining and this ratio of 25 is
expected to go down to almost five soon with the introduction of advanced materials and innovative systems [23].
In photocatalysis, photonic energy is converted directly to chemical energy (i.e., hydrogen) without the need to have electricity
as an intermediate medium. The amount of photonic energy is directly proportionate to the frequency of the incoming photon.
This relationship is generally stated by hn where h is the Planck’s constant and n is the frequency of the incoming photon. When an
incoming photon touches the surface of the photocatalyst, an electron–hole pair is produced. As a result, electrical charge is
obtained by splitting this electron–hole pair and this charge is used to split water into hydrogen and oxygen. For this process to
proceed from photon to electron–hole pair and to hydrogen, the photocatalyst must have a suitable band gap energy. In addition,
the photocatalyst must have appropriately positioned conduction and valance bands for oxidation and reduction reactions.
Additionally, prompt production and splitting of electron–hole pairs is important in photocatalytic hydrogen production.
Therefore the photocatalyst must be selected appropriately. In the literature, semiconductors (such as TiO2) and metal oxides
(such as Fe2O3) are comprehensively investigated as photoactive materials. Correspondingly, chemically designed and engineered
complex supramolecular substances are used in photocatalytic hydrogen production reactions. Acar et al. [24] have comparatively
investigated numerous simple and complex photocatalysts with respect to their H2 generation amounts, efficiencies, and impacts
on health and the environment.
Acar et al. [25] have reviewed the progress and performance of photocatalytic hydrogen production methods for better
sustainability by considering the following photoreduction and photooxidation reactions:

hn
Photoreduction : 2H2 O þ 2e ⟶ H2 þ 2OH ð23Þ

hn
Photooxidation : 2H2 O ⟶ O2 þ 4Hþ þ 4e ð24Þ

PEC cells convert solar energy to an energy carrier via light stimulated electrochemical processes. In a PEC, solar light is
absorbed by one or both of the photoelectrodes and at least one of them is a semiconductor. PECs can produce either chemical or
electrical energy. They are also used to treat hazardous aqueous wastes [26]. The operating procedure of the semiconductor in a
PEC is analogous to a PV panel. In both systems, photons with higher energy than the band gap of the photoactive material
produce electron–hole pairs and this resulting electricity is utilized to oxidize and reduce water. The PEC method combines solar
energy processing and water electrolysis in one single reactor. This is the main benefit of the PEC method. In PEC, a separate
electricity production unit, such as a PV cell is not required and consequently PEC requires less space. In the literature, there are
numerous types of photoactive materials considered for use in PEC. So far, TiO2 is considered as the most promising photoactive
material to be used in PEC. Along with TiO2, numerous other semiconductors have been investigated, for example, ZnO, Fe2O3,
BiVO4, and WO3. In addition, metal nitrides and phosphides (such as Ta3N5 and GaP), metal oxynitrides (e.g., TaON), and n- and
p-type silicon have been considered in the literature by many researchers. Rabbani et al. [27] have integrated PEC with chloralkali
reactors and demonstrated the system performance in batch type operation. Acar and Dincer [28] have integrated these studies and
enhanced system performance of PEC and chloralkali reactors by using an integrated continuous type hybrid hydrogen production
process.

3.1.3.7 Dark Fermentation


Biochemical energy is the energy stored in organic and living matter. This energy can potential to be harvested by living beings to
generate hydrogen in either the absence or presence of photonic energy. This photonic energy can come from the sun or a source
mimicking the sun. If this biochemical to hydrogen energy conversion occurs in the absence of any light, it is called dark
fermentation. Dark fermentation can also take place when the light supply is reduced below a certain level.
Dark fermentation reactors are simpler and more affordable than the photofermentation reactors as dark fermentation does not
need any photon energy processing components. There are numerous other advantages of hydrogen production via dark fer-
mentation. For example, hydrogen production ability from organic waste helps to regulate, stabilize, and eliminate waste, which
can potentially eliminate any risk of biological waste pollution. As another example, dark fermentation can be combined with
wastewater management facilities to generate H2 from wastewater.
Hydrogen production from organic and biological waste and wastewater can potentially lower hydrogen production costs as
the process input is waste of other processes so it is affordable, abundant, and easily available. This approach also lowers the
operating cost of waste treatment facilities.
Koutrouli et al. [29] have investigated hydrogen production from the waste stream of olive oil production facilities and
demonstrated a hydrogen production up to 640 g/t of olive pulp.
Das and Veziroglu [30] have studied hydrogen production via dark fermentation of glucose and shown a hydrogen production
amount around 77 g/kg of glucose. Some of the main issues related to hydrogen production via dark fermentation include the low
generation capacities per unit of the capital investment.
Benefits, drawbacks, and future predictions of dark fermentation, biophotolysis, and photofermentation-based hydrogen
production are summarized in Table 6.
Hydrogen Production 15

Table 6 Benefits, drawbacks, and future predictions of dark fermentation, biophotolysis, and photofermentation

Process Benefits Drawbacks Future predictions

Biophotolysis Abundant material resources (water) H2 and O2 separation Short-term solutions available
Carbon free process Low system efficiency Immobilization and improvement
potential
No other products than H2 and O2 Big surface areas needed Advancements in materials science
Photofermentation Availability of material resources Low volumetric H2 generation Metabolic engineering requirement
(waste streams)
Almost complete conversion reaction Low system efficiency Short-term solutions available
Big surface areas needed Advancements in materials science
Dark fermentation Supply availability (waste streams) Byproduct generation Metabolic engineering requirement
Simple reactor Lack of consistency Two stage systems to lower chemical
oxygen demand (COD)
High generation yields COD removal issues

Source: Reproduced from Hallenbeck PC, Abo-Hashesh M, Ghosh D. Strategies for improving biological hydrogen production. Bioresour Technol 2012;110:1–9.

3.1.3.8 High Temperature Electrolysis


HTE is a method of electrolysis where steam is dissociated to H2 and O2 at temperatures between 700 and 10001C. In electrolysis,
system efficiencies increase with increasing operating temperatures. Therefore HTE is known to be more efficient than the con-
ventional electrolysis at room temperature. As a result, the electrical energy requirement of HTE is lower than the electrical energy
requirement of conventional electrolysis.
The first step of HTE is heating up the water to first evaporate and then to operating temperatures, which requires significant
thermal energy input. This thermal energy can either be supplied directly by direct injection of steam to the electrolyzer or
indirectly by using an external heat source. If this heat source is a clean one, such as geothermal, solar, or recovered heat, HTE
produces hydrogen with nearly zero GHG emissions.
In this method, the system components must have specific properties in order to withstand the operating conditions of HTE.
This brings some challenges when developing high-temperature electrolyzers; some of these challenges are: (1) chemical and
physical stability at high operating temperatures, ionic concentrations and/or low electrical conductivities; (2) porous and che-
mically stable electrode requirements in extremely reducing and oxidizing environments with enhanced electrical and thermal
conductivities; and (3) engineering chemically and physically stable materials at high operating temperatures and extremely
reducing and oxidizing mediums.

3.1.3.9 Hybrid Thermochemical Cycles


Hybrid thermochemical cycles’ operating temperatures are significantly lower with respect to thermochemical water splitting cycles
discussed in the previous sections. Thermal and electrical energy sources are used to meet the external energy requirements of each
reaction in the thermochemical water splitting cycles. Hybrid thermochemical cycles’ advantage over conventional thermo-
chemical cycles is their lower quality heat input requirement. This means hybrid cycles use heat at lower temperatures, which can
come from a variety of sustainable processes as recovered heat.
Among the hybrid thermochemical cycles investigated in the literature, Cu–Cl cycle is an outstanding one with numerous
advantages. Cu–Cl has reasonably low operating temperatures that do not surpass 5501C. And among the various available types
of Cu–Cl cycles, the five step Cu–Cl process is the most heavily investigated one in the literature. In five step Cu–Cl cycle, there are
three endothermic reactions, one electrochemical reaction, and one drying process. In Cu–Cl cycles, part of the external heat
supply is used to directly support the endothermic cyclic reactions and the rest is used to produce power, which is also required by
the hybrid thermochemical cycle itself. The main benefit of the Cu–Cl cycles is hydrogen production from low-temperature heat
supplies. This heat can be harvested from sustainable energy sources, which makes the Cu–Cl cycle itself very sustainable as well.
This low-temperature heat can be recovered sustainably from power plants, various industries, concentrated solar processes, and
municipal waste incineration.

3.1.3.10 Coal Gasification


Coal is currently the most financially feasible and technically practical energy source for large-scale hydrogen production due to the
present technological state of coal processing and global coal reserves. In coal gasification, coal is partially oxidized by using steam
and O2 in a high-temperature and high-pressure reactor. The products of the first step of coal gasification are predominantly H2
and CO along with some steam and CO2. This product gas mixture is called syngas. Then, the syngas is sent to a unit to be
processed in a shift reaction aiming to enhance the hydrogen amount in the syngas. Later, the product gas is sent to a treatment
unit if there is a need to eliminate elemental sulfur or sulfuric acid in the product mixture. Additionally, this syngas can further be
utilized for power generation via turbines.
16 Hydrogen Production

Coal gasification-based hydrogen production has several advantages over the other alternative hydrogen production methods,
such as electrolysis. However, despite these advantages, because of the high carbon content of coal, coal gasification has very high
CO2 emissions, higher than any of the alternative hydrogen production methods. Therefore coal gasification requires the use of
CCS technologies to eliminate the GHG emissions.
At present, the cost of hydrogen via coal gasification is to some extent more expensive than that of SMR. However, it should be
noted that the cost components of hydrogen production from coal is different than that of fossil fuels. In coal fueled systems, the
raw materials are cheaper but the unit capital costs are expensive [31].

3.1.3.11 Fossil Fuel Reforming


Some examples of hydrogen production via fossil fuel reforming are steam reforming, POX, and ATR. The benefits and drawbacks
of these methods are summarized in Table 7. Similar to coal gasification, hydrogen production via fossil fuel reforming has
byproducts of CO and CO2.
Steam reforming does not require oxygen for the hydrogen production reaction. However, steam reforming does usually
require an external heat source. The operating temperature of steam reforming is lower than that of POX and ATR. In addition,
steam reforming has higher H2/CO ratios.
In hydrogen production via POX (combustion), fossil fuels are partially oxidized by using oxygen. Heat source of this reaction
is recovered from the partial combustion reaction itself. No catalyst is required in POX and this process is more tolerant to sulfur
compared to steam and ATR.
Hydrogen production via ATR has lower operating pressures compared to POX. Both ATR and POX have no need for an
external heat supply. On the other hand, both of these methods need pure oxygen feed. This makes these processes more complex
and expensive since additional oxygen separation and processing units are required. Compared to the selected fossil fuel reforming
technologies, steam reforming (predominantly SMR) is the cheapest and most commonly used one for hydrogen production.

3.1.3.12 Biophotolysis and Photofermentation


Biophotolysis and photofermentation are biochemical reactions driven by photonic energy to produce hydrogen by using water as
the material resource. Kotay and Das [32] have classified biophotolysis-based hydrogen production into three categories: direct,
indirect, and photofermentation. In biophotolysis, various photosensitive microorganisms are utilized as biochemical conversion
devices. These microorganisms are placed in specifically developed photobioreactors. Amongst the alternative microorganisms
studied in the literature, the most appropriate and efficient ones are microalgae. This is because of the fact that microalgae can be
cultured and they can produce hydrogen in closed systems. As a result, capturing the product hydrogen is easier with microalgae.
The main benefit of biophotolysis is the capability to generate hydrogen from water in mild environments such as at moderate
temperatures and pressures (e.g., standard temperature and pressure). On the other hand, biophotolysis is only confirmed at the
laboratory scale and it has to be fully commercialized first before its market introduction. The overall hydrogen production
reactions in biophotolysis and photofermentation can be written as

hn
6H2 O þ 6CO2 ⟶ C6 H12 O6 þ 6O2 ð25Þ

hn
C6 H12 O6 þ6H2 O ⟶ 6CO2 þ12H2 ð26Þ

3.1.3.13 Artificial Photosynthesis


Artificial photosynthesis is a process that mimics a very famous biochemical reaction: natural photosynthesis. The artificial
photosynthesis system includes an enzyme bed reactor to fix CO2 in the air (or any other source needing CO2 to be removed). This

Table 7 Benefits and drawbacks of selected fossil fuel reforming methods for hydrogen production

Technology Benefits Drawbacks

Steam reforming Highly industrial process Highest greenhouse gas (GHG) emissions
No oxygen input need
Lowest operating temperatures
Highest H2/CO ratio
Autothermal reforming Operating temperature lower than partial oxidation (POX) Limited commercial experience
Less methane slip Air/oxygen requirement
POX Fewer desulfurization needs Low H2/CO ratio
No catalyst requirement High operating temperatures
Less methane slip Complex handling process

Source: Reproduced from Holladay JD, Hu J, King DL, Wang Y. An overview of hydrogen production technologies. Catal Today 2009;139(4):244–60.
Hydrogen Production 17

reactor is fueled by hydrogen energy and bioelectric transducers. The key components of an overall artificial photosynthesis system
are:

• Electricity production via PV panels to meet any electrical energy needs of the systems.
• Dry agriculture: carbohydrates (food), liquid fuels, chemical feedstocks, and polymers for fiber manufacture can be produced
with near or absolutely minimum water requirements. Artificial photosynthesis can lower conventional agricultural water use
by almost thousands of times.
• Hydrogen production: electrochemical water dissociation into H2 and O2 is accomplished by mimicking photosynthesis.
Even though artificial photosynthesis technology is in its early stages and cannot yet be used in large-scale plants, it has a
substantial prospective to reduce global water demand and substantiate clean energy systems by producing power and hydrogen
from photonic energy directly.

3.1.3.14 Photoelectrolysis
In photoelectrolysis, heterogeneous photocatalysts are coated either on one or both of the electrodes to enhance hydrogen
production. In photoelectrolysis, hydrogen production occurs both directly from photonic energy and via the electrical energy
coming from the electrodes. For that reason, in photoelectrolysis, both photonic and electrical energies are used to generate
chemical energy (i.e., hydrogen). There are five steps in the photoelectrolytic hydrogen production process: (1) electron–hole pair
production with the incoming photon with higher energy than the band gap of the photocatalyst; (2) electron flow from the anode
to the cathode, which generates electrical current; (3) water dissociation into H þ ions and gaseous O2; (4) H þ ions reduction at
the cathode generating gaseous H2; and (5) product gas separation, treatment, and storage. Photoelectrolysis efficiency is strongly
influenced by the type, crystalline structure, and surface properties of the photoactive substance and corrosion resistance and
reactivity of all system components. Generally, there is a compromise concerning photoelectrode stability and photon to hydrogen
energy conversion efficiency. The problem is most of the photoactive materials with high conversion efficiencies usually are less
stable in electrolytic mediums. In the meantime, the chemically and physically stable photoactive materials have low system
efficiencies [33].

3.1.3.15 Summary
Some critical challenges, key research and development requirements, and major advantages of the hydrogen production methods
discussed in this section are summarized in Table 8.
At the present technological advancement level, natural gas fueled hydrogen generation in large scales is the most affordable
hydrogen production method available in the literature. In all energy systems, supply security is a critical requirement. Therefore
energy and material supply security must be considered in all hydrogen production systems. Reaching optimized capital, oper-
ating, and maintenance costs; reaching high energy conversion and overall system efficiencies; maintaining lower impurity and
emission levels; and enhancing the active contribution of renewable energy resources are the main challenges to be addressed
when developing sustainable hydrogen production systems.
At the end of the day, the goal of hydrogen production is to have sustainable, clean, affordable, and reliable energy systems.
Therefore hydrogen has to be generated by using clean energy resources rather than fossil fuels. Production yields and costs, system
efficiencies, overall system reliabilities, and any negative environmental impacts are the primary focus areas of research in
sustainable hydrogen production. It is very well known that accomplishing carbon free energy systems is very highly unlikely
without hydrogen energy systems. Therefore this chapter discusses and comparatively evaluates some key hydrogen production
methods to find the ones with lowest possible production cost, environmental and social impact, along with highest possible
system efficiency. With this chapter, the aim is to tackle the challenges related to negative impacts of too much fossil fuel use with
sustainable hydrogen production systems.

3.1.4 Efficiency Evaluation of Hydrogen Production Methods

The most general definition of energy and exergy efficiencies is the ratio of the energy or exergy content of the desired product to
that of the material and/or energy resource. More specifically, energy efficiency of a hydrogen production method is calculated by
the following equation:
_
mLHV H2
Z¼ ð27Þ
E_ in
where m_ is the mass flow rate of produced hydrogen, LHV is the lower heating value of hydrogen (121 MJ/kg), and E_ in is the rate of
energy input to the process. The following equation is used for exergy efficiency:
_ ch
mex H2
c¼ ð28Þ
_ in
Ex
here, exch _
H2 is the chemical exergy of hydrogen and Exin is the exergy input rate supplied to the hydrogen production process.
18
Hydrogen Production
Table 8 Critical challenges, key research and development requirements, and major advantages of the hydrogen production methods discussed in this section

Fossil fuel reforming Biofuel reforming Coal and biomass Thermochemical method Water electrolysis Photoelectrochemical Biological method
gasification (PEC) method

Critical challenges
Capital costs Capital costs Reactor costs Reactor costs Low efficiency Low efficiency Low efficiency
Design Operation and Low efficiency Technological level Capital costs Photoactive material Microorganism
maintenance costs efficiency functionality
Operation and Design Feedstock quality Efficient and System integration Reactor costs Reactor material selection
maintenance costs Feedstock impurities Carbon capture and stable substances Design problems Technological level Technological level
storage (CCS)

Main research and development requirements


Efficiency and cost Hydrogen yield and Low cost and efficient Robust, low cost materials Durable and cheap Durable and efficient Microorganism
efficiency purification materials photocatalyst functionality
Low cost and efficient Low-temperature Co-fed gasifiers Ease of manufacture and Corrosive-resistant Low cost materials New organisms
purification production application membranes
Feedstock pretreatment Low cost and efficient CCS System optimization Durable, active, and cheap Active, stable, and cheap Inexpensive methods
purification catalysts supporting materials
Optimization Optimization Hydrogen quality High volume, low cost, Large-scale applications High volume production Low cost and durable
flexible system design material
Automated process Regional best feedstock Cost of feedstock Efficient heat transfer Storage and production System control System optimization
control preparation rate
Reliability Feedstock pretreatment Tolerance for impurities Reliability Reliability Power losses High capacity and low cost
systems
Major advantages
Most viable approach Viability Low cost syngas Clean and sustainable No pollution with Low operation Clean and sustainable
production renewable energy temperature
sources
Lowest current cost Existing infrastructure Abundant and cheap Recycled chemicals Existing infrastructure Clean and sustainable Tolerant of diverse water
feedstock conditions
Existing infrastructure Integration with fuel cells Self sustaining

Source: Reproduced from Dincer I, Acar C. Review and evaluation of hydrogen production methods for better sustainability. Int J Hydrog Energy 2015;40(34):11094–111.
Hydrogen Production 19

In addition to energy and exergy efficiency, hydrogen production performance is evaluated by various additional criteria.
Although among the available ones in the literature, energy efficiency is one of the most commonly used ones. In solar hydrogen
production, if every photogenerated electron and hole are assumed to be utilized in the water dissociation reaction, the following
equation can be used to calculate the overall STH efficiency (ZSTH) as
_ out  W
W _ in jphoto ðVredox  Vbias Þ
ZSTH ¼ ¼ ð29Þ
I_ I_
here, I_ is the power input, which is solar irradiation, W_ out and W _ in are electrical power output and input, respectively. Vredox is
typically accepted to be 1.23 V (at around 298K). This amount is water dissociation reaction’s Gibbs free energy change which is
237 kJ/mol. On the other hand, the enthalpy change of water splitting reaction, which is 286 kJ/mol, is occasionally utilized. This
gives a redox potential of 1.48 V. The potential calculated based on the enthalpy change is applicable in cases where the generated
hydrogen is planned to be burned in a combustion chamber. While the potential calculated based on the Gibbs free energy change
is more suitable in cases where the hydrogen is transformed into electricity in a fuel cell. In cases where the intended utilization of
product hydrogen is not known, it is most likely more suitable to take the Gibbs free energy change as basis as it gives lower
efficiencies. When AM1.5G solar irradiation (I_ ¼ 1000 W/m2) is used and no bias is applied (Vbias ¼ 0 V), overall STH efficiency can
possibly be calculated straight from the photocurrent (jphoto):
ZSTH ¼ 1:23  jphoto ð30Þ
where unit of ZSTH is % and jphoto is mA/cm . Hence, a photocurrent of around 8 mA/cm is necessitated to accomplish the 10%
2 2

overall STH efficiency objective for financially sustainable hydrogen production.


It should be emphasized that Vbias in Eq. (29) denotes the definite potential difference amongst the working and counter
electrodes. This potential difference can be indirectly determined by using a three-electrode measurement technique. In this
method, the working electrode potential is measured comparatively with respect to the reference electrode potential. Since this
method commonly causes some mistakes, it would be better to avoid using the three-electrode measurement technique if possible.
Measuring the quantity of hydrogen production by mass spectrometry and/or gas chromatography is a more direct technique to
evaluate the STH efficiency. Water displacement in an inverted buret is another way to verify the amount of production. With this
method, the STH efficiency is calculated by
n_ H 2 G1f ;H 2
ZSTH ¼ ð31Þ
I_
where n_ H2 is the hydrogen generation rate at the irradiated surface (mol/s/m2) and G1f ; H2 is the Gibbs free energy change of
hydrogen production (237 kJ/mol).
When reporting STH efficiencies, the following assumptions are made: (1) the reaction is stoichiometric (H2:O2 molar ratio is
2:1); (2) there are no side reactions happening as a result of additional chemicals, such as electrolytes, electron donors, and/or
acceptors; and (3) the irradiation source is stated appropriately and it sufficiently agrees with the AM1.5G spectrum in con-
centration and spectral distribution.
In cases where efficiency restraining criteria are being identified, it is an advantageous approach to use the quantum efficiency as
a function of wavelength. External quantum efficiency is defined as the ratio of the incoming photons, which are transformed to
electrons to the external circuit current measurement. This term is called the incident photon-to-current conversion efficiency
(IPCE), which is calculated as
 
hc jphoto ðlÞ
IPCEðlÞ ¼ ð32Þ
e _
lIðlÞ
In the literature, there are IPCE values reported to be higher than 80% for the photooxidation of water, for instance, for WO3
and for TiO2 under UV irradiation. An additional beneficial efficiency indicator is the absorbed photon-to-current conversion
efficiency (APCE). Compared to the IPCE, the APCE accounts for the reflection losses as well. APCE is frequently mentioned as the
internal quantum efficiency, and it is calculated based on the IPCE, which is shown in the following equation:
IPCEðlÞ IPCEðlÞ
APCEðlÞ ¼ ¼ ð33Þ
AðlÞ 1RT
here, A, R, and T are the optical absorption, reflection, and transmission, respectively. The optical absorption (A) should not be
mistaken as the absorbance or optical density (OD). OD is specified by:
 
T aL
OD ¼  Log ¼ ð34Þ
1R 2:303
where a is the absorption coefficient and L is the substrate thickness. APCE should be used in cases where it is aimed to estimate
electron–hole recombination inside the semiconductor. The IPCE is more applicable to utilize when evaluating conversion
efficiencies of overall devices and systems. In some special cases, the IPCE is evaluated in monochromatic illumination to estimate
the photocurrent in real solar light:
Z
Jsolar ¼ ðIPCEðlÞ  FðlÞ  eÞdl ð35Þ
20 Hydrogen Production

In this equation, Jsolar is the overall solar photocurrent in A/m2 and F(l) is the photon flux of solar irradiation in
photons/m2/s. The photon flux might be determined based on the tabulated solar irradiation data, E(l), via
EðlÞ
FðlÞ ¼ ð36Þ
hc=l
An important underlying assumption for Eq. (35) is that there is generally a direct connection amongst the monochromatic
photocurrent density and the light intensity; however, there are exceptions. A superlinear photocurrent escalation with increasing
light intensity is occasionally detected for nanostructured substances having above average concentrations of moderately narrow
surface or interface traps. In such cases, these interface traps firstly require to be filled up before the charge carriers could make it to
the electrolyte or the rear contact. On the other hand, slow charge transportation or charge transference through the substrate
interface could cause a build-up of free charge carriers at or close to the interface. As a result, the amount of recombination
proliferates and this will cause a sublinear intensification of the photocurrent with light intensity.

3.1.5 Comparative Evaluation of Hydrogen Production Methods

3.1.5.1 Comparison of Environmental Impact


CO2 emissions are regarded as the most important sources of GHG emissions. Increasing CO2 emissions have substantially
negative effects on the environment and health of all living beings. At this time, there are various available solutions to reduce CO2
emissions from energy systems. Some of these solutions are carbon capture and sequestration (CCS), CO2 treatment as an input in
a different industry, such as methanol production, etc. More detailed information on CO2 emissions and how to eliminate or
minimize them is provided in Refs. [34,35]. Carbon neutral energy systems with energy and material supply security are exten-
sively investigated in the literature. Hydrogen has to be produced by using clean and sustainable sources and energy systems and
renewable materials to reach the goal of lowering CO2 emissions of energy systems.
An extensive evaluation of CO2 emissions of a process requires a life cycle assessment (LCA) approach. Several available LCA
procedures developed based on the ISO standards have been summarized and provided by the Center of Environmental Science of
Leiden University in its “Operational Guide to the ISO Standards” [36]. The environmental impact types taken into account in this
section are taken from this operational guide. Global warming potential (GWP) and acidification potential (AP) are selected to
assess the environmental impact of the selected hydrogen production methods. GWP (kg CO2 eq.) indicates the amount of CO2
emissions. AP (g SO2 eq.) refers to SO2 discharge on soil and into water and measures the change in degree of acidity [37].
In this chapter, the GWP and AP LCA results previously provided by Bhandari et al. [38] and Ozbilen et al. [39] are selected as
the basis of environmental impact comparison of selected hydrogen production methods. LCA has usually four major steps: (1)
defining the goal and scope, (2) collection of the life cycle inventory on all material and energy flows during the entire life cycle of
the process, (3) life cycle impact assessment (LCIA) built on the inventory results by evaluating the importance of each item, and
(4) evaluation of LCIA results and discussion of methods to minimize or eliminate the negative environmental impacts. The
environmental impact comparison of selected hydrogen production methods, in terms of GWP and AP, are presented in Fig. 4. The
fossil fuel-based hydrogen production methods (coal gasification, fossil fuel reforming, and plasma arc decomposition) are seen to

14 40
GWP
Global warming potential (kg CO2 / kg H2)

AP
Acidification potential (g SO2 /kg H2)

12 35

30
10
25
8
20
6
15
4
10

2 5

0 0
M1
M2
M3
M4
M5
M6
M7
M8
M9
M10
M11
M12
M13
M14
M15
M16
M17
M18
M19

Methods
Fig. 4 Comparison of global warming potential (GWP) and acidification potential (AP) of the selected hydrogen production options (per kg of
hydrogen).
Hydrogen Production 21

2.0
1.8

Social cost of carbon ($/kg H2)


1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
M1
M2
M3
M4
M5
M6
M7
M8
M9
M10
M11
M12
M13
M14
M15
M16
M17
M18
M19
Methods
Fig. 5 Social cost of carbon (SCC) comparison of the selected hydrogen production options (per kg of hydrogen).

be the most environmentally harmful methods. Although it has relatively low GWP, the AP of biomass gasification is the highest
compared to the other selected methods. From Fig. 4, it can also be seen that photonic energy and hybrid thermochemical cycle-
based hydrogen production are the most environmentally benign of the selected methods, in terms of CO2 emissions and APs.

3.1.5.2 Comparison of Social Cost of Carbon


The negative impacts of GHG emissions on the environment and health of all living beings have been one of the commonly
accepted critical issues. SCC is a term that is used to relate the negative impacts of CO2 emissions to their corresponding economic
costs. SCC can be defined as the marginal external cost of a unit of CO2 emissions. An integrated assessment framework (IAF) is
used to estimate the SCC of a process or product. In this framework, a reference socioeconomic scenario is used with a model. The
model relates CO2 emissions to temperature change and then to economic damages. These economic damages also take external
costs of CO2 emissions into account. Some examples of the external costs are: increased healthcare costs, insurance premiums,
lower productivities, etc.
In the first step of SCC evaluation, the baseline socioeconomic scenario is defined by the state of the evaluated technology with
its production capacity, rate, and emissions and the population. The relationship between the GHG concentrations and tem-
perature deviations from seasonal and historic averages is used to evaluate the impacts of climate change. Next, the deviations
from the reference socioeconomic scenario and the impact on the economy are calculated. After that, the reference scenarios are
marginally perturbed by adding or removing marginal units of CO2 emissions. Social wellbeing is generally determined by the
amount of consumption and the choice of discounting parameters. In SCC analysis, social welfare is determined for each reference
and marginally perturbed scenario. The normalized variation in anticipated wellbeing between the reference and marginally
perturbed scenarios is regarded as the SCC [40].
In this chapter, the SCC of selected hydrogen production alternatives is determined based on the results provided by Kopp and
Mignone [40]. The SCC of each alternative is calculated by taking SCC to be $160/t of CO2 emissions. Fig. 5 represents the SCC
evaluation findings of the selected hydrogen production alternatives. Here, it can be seen that photonic energy and hybrid
thermochemical cycle-based hydrogen production alternatives have the lowest SCC, which can be also interpreted as the most
advantageous. Fossil fuel powered hydrogen production alternatives (i.e., coal gasification, fossil fuel reforming, plasma arc
decomposition, etc.) have the highest SSC, due to their very high and mostly harmful CO2 emissions.

3.1.5.3 Comparison of Hydrogen Production Cost


There are numerous uncertainties related to determining the true cost of hydrogen production. This is mainly due to the fact that
the production cost is strongly influenced by the technological development state, whether or not a well-developed infrastructure
exists, and the cost of energy and material resources.
Fig. 6 presents the average costs of selected hydrogen production methods, based on the findings stated earlier in peer reviewed
publications. Amongst the selected options, the average hydrogen production costs of water electrolysis, thermochemical water
splitting, biomass gasification, photocatalysis, coal gasification, and fossil fuel reforming are obtained from Parthasarathy and
Narayanan [41]. The average hydrogen production costs of plasma arc decomposition, thermochemical biomass conversion and
reforming, dark fermentation, biophotolysis, photofermentation, artificial photosynthesis, and photoelectrolysis are taken from
Uddin et al. [42]. The average hydrogen production costs of thermolysis, PV-based electrolysis, HTE, and hybrid thermochemical
22 Hydrogen Production

12.0

10.0

Production cost ($/kg H2)


8.0

6.0

4.0

2.0

0.0
M1
M2
M3
M4
M5
M6
M7
M8
M9
M10
M11
M12
M13
M14
M15
M16
M17
M18
M19
Methods
Fig. 6 Production cost comparison of selected hydrogen production options (per kg of hydrogen).

90
Energy
80
Exergy
70
Efficiency (%)

60
50
40
30
20
10
0
M1
M2
M3
M4
M5
M6
M7
M8
M9
M10
M11
M12
M13
M14
M15
M16
M17
M18
M19
Methods
Fig. 7 Comparison of energy and exergy efficiencies of selected hydrogen production options.

cycles are compiled from Ngoh and Njomo [43]. And lastly, the average hydrogen production cost of PEC cells is acquired from
Trainham et al. [44].
Fig. 6 shows that fossil fuel powered hydrogen production methods (SMR, coal and biomass gasification, and plasma arc
decomposition) are the cheapest hydrogen production options among the selected alternatives. These options are followed by
thermochemical cycles and biomass-fueled hydrogen production methods, and hybrid thermochemical cycles, which have
potential to become cost competitive with fossil fuel powered methods in the near future. On the other hand, PEC cells are the
most expensive hydrogen production methods among the selected options. In spite of this, it should be remembered that PEC
technology is mostly in its primary research and development state so its efficiency construction, operation, and maintenance costs
are expected to decrease as innovative technologies and advanced materials are developed. One of the most important benefits of
PECs is their ability to be employed locally; this is particularly advantageous in geographically remote locations with no or limited
access to the grid.

3.1.5.4 Comparison of Energy and Exergy Efficiencies


The energy and exergy efficiencies of the selected hydrogen production methods are obtained from Holladay et al. [31], Ismail and
Bahnemann [45], Singh and Wahid [46], Ibrahim et al. [47], Bicakova and Straka [48], and Dincer and Zamfirescu [49] and these
results are presented in Fig. 6. From Fig. 7, it can be understood that fossil fuel powered hydrogen production options (such as
reforming, plasma arc decomposition, and coal gasification) and biomass gasification have higher energy and exergy efficiencies
than the selected options. Alternatively, photonic energy powered hydrogen production technologies have the lowest energy and
exergy efficiencies compared to the rest of the evaluated production options. As discussed earlier, this is mainly because of the fact
Hydrogen Production 23

Table 9 Normalized ranking comparison of selected hydrogen production options

Method Energy Exergy Cost Social cost Global warming Acidification


efficiency efficiency of carbon (SCC) potential (GWP) potential (AP)

M1 Electrolysis 5.30 2.50 7.34 3.33 3.33 8.86


M2 Plasma arc decomposition 7.00 3.20 9.18 0.83 0.83 5.14
M3 Thermolysis 5.00 4.00 6.12 7.50 7.50 7.43
M4 Thermochemical water splitting 4.20 3.00 8.06 9.17 9.17 9.43
M5 Biomass conversion 5.60 4.50 8.10 6.67 6.67 2.00
M6 Biomass gasification 6.50 6.00 8.25 5.83 5.83 0.00
M7 Biomass reforming 3.90 2.80 7.93 6.25 6.25 0.86
M8 Photovoltaic (PV) electrolysis 1.24 0.70 4.50 7.50 7.50 7.71
M9 Photocatalysis 0.20 0.10 5.19 9.58 9.58 9.71
M10 Photoelectrochemical (PEC) method 0.70 0.15 0.00 9.58 9.58 9.71
M11 Dark fermentation 1.30 1.10 7.52 9.58 9.58 9.71
M12 High-temperature electrolysis 2.90 2.60 5.54 7.92 7.92 8.57
M13 Hybrid thermochemical cycles 5.30 4.80 7.41 9.43 9.43 9.02
M14 Coal gasification 6.30 4.60 9.11 0.00 0.00 1.31
M15 Fossil fuel reforming 8.30 4.60 9.28 2.50 2.50 5.71
M16 Biophotolysis 1.40 1.30 7.27 7.50 7.50 9.71
M17 Photofermentation 1.50 1.40 7.61 9.58 9.58 9.71
M18 Artificial photosynthesis 0.90 0.80 7.54 9.58 9.58 9.71
M19 Photoelectrolysis 0.78 0.34 7.09 8.33 8.33 9.71
Ideal Zero emissions and cost, 100% efficiency 10.00 10.00 10.00 10.00 10.00 10.00

that photonic energy-based hydrogen production is in its early research and development phase. Therefore these efficiencies are
expected to increase with the introduction of innovative novel technologies and advanced materials.

3.1.5.5 Overall Comparison


In this section, the environmental, social, economic, and efficiency evaluation findings are normalized with the intention of
comparing the selected hydrogen production options in an effective manner. Normalization of GWP, AP, SCC, and hydrogen
production costs are conducted based on the subsequent formula:
Maximum  Method i
Rank ðMethod iÞ ¼  100 ð37Þ
Maximum
The ranking is between 0 and 10, where 0 means poor performance and 10 indicates the ideal case (zero cost and zero
emissions). Lower costs and emissions are given higher rankings. “0” is assigned to the highest cost and emissions in selected
categories. For example, in terms of GWP, the coal gasification method gives the highest emissions; therefore, the GWP ranking of
coal gasification is assigned to be “0.” Efficiencies are normalized based on the following equation:
Efficiency rank ðMethod iÞ ¼ EfficiencyðMethod iÞ  10 ð38Þ
The ranking range is again between 0 and 10; 0 means poor performance and 10 indicates the ideal case (100% efficiency).
Higher rankings mean higher efficiencies. The normalized emissions, cost, and efficiency rankings are presented in Table 9. The
theoretical ideal hydrogen production option has zero production cost, no harmful emissions, and also zero SCC. The energy and
exergy efficiencies of this ideal hydrogen production method are 100%. In terms of energy and exergy efficiencies, fossil fuel and
biomass powered options have closest to ideal performance. On the other hand, biomass-based methods have significantly high
AP (stated as lowest AP rankings) and relatively high GWP and SCC compared to the selected options. Despite their significantly
low emissions, the low average normalized rankings of photonic-based hydrogen production methods are due to their low system
efficiencies and production costs.

3.1.6 Comparative Evaluation Results of Hydrogen Production Methods

3.1.6.1 Hydrogen Production From Electrical Energy


The electrical energy-based hydrogen production methods evaluated in this chapter are electrolysis and plasma arc decomposition.
Electrical energy-based hydrogen production options have higher energy and exergy efficiencies compared to the other selected
alternatives investigated in this chapter. Hydrogen production cost results show that electrical energy-based hydrogen production
methods are cost competitive to fossil fuel-based methods. And it should be noted that unlike the electrical-based options, fossil
fuel powered hydrogen production methods are more mature in terms of technological advancement levels. On the other hand,
emissions results indicate high GWP and AP, which shows that electricity should come from clean energy sources in order to make
24 Hydrogen Production

Energy
efficiency
10
9
8
7
6
AP Exergy
5
efficiency
4
3
2
Electrolysis
1
0 Plasma arc
decomposition
Ideal

GWP Cost

SCC
Fig. 8 Normalized ranking comparison of electrical-based hydrogen production methods. AP, acidification potential; GWP, global warming
potential.

electrical-based hydrogen production more environmentally benign. The normalized rankings of electrolysis and plasma arc
decomposition are presented in Fig. 8. Electrolysis is closer to the ideal case than plasma arc decomposition in terms of GWP, AP,
and SCC. This is because electrolysis has less CO2 and SO2 emissions and lower SCC. Conversely, plasma arc decomposition has
higher efficiencies and lower production costs. Overall, average normalized ranking of electrolysis is higher than that of plasma arc
decomposition. There is a substantial focus on research and development of electrolysis in the literature. When powered by
renewable energies, electrolysis can potentially provide hydrogen with almost zero harmful emissions. In addition, innovative
technologies and materials can enhance efficiencies and costs of electrolysis.

3.1.6.2 Hydrogen Production From Thermal Energy


Thermolysis, thermochemical water splitting, thermochemical conversion of biomass, gasification, and reforming are selected
hydrogen production options powered by thermal energy. The normalized ranking comparison of thermal energy powered
hydrogen production options is presented in Fig. 9. These options have significantly high energy and exergy efficiencies and close
to ideal GWP, SCC, and hydrogen production costs compared to other alternatives discussed in here so far. On the other hand,
thermal energy-based hydrogen production options, particularly biomass gasification, have considerably high SO2 emissions
(indicated as high AP). Amongst the thermal energy powered hydrogen production alternatives, biomass gasification has the
highest energy and exergy efficiencies and lowest hydrogen production cost. Thermochemical water splitting has the lowest GWP,
AP, and SCC. Biofuel reforming has the lowest energy and exergy efficiencies. Thermolysis has the highest hydrogen production
cost. And the least environmentally benign thermal energy powered hydrogen production method amongst the selected options is
biomass gasification. Amongst the selected thermal energy powered hydrogen production options, thermochemical water splitting
has closest to ideal case performance with highest average normalized rankings. On the other hand, biomass (biofuel) reforming
has the lowest average normalized rankings (Table 9). Thermochemical water splitting has relatively high normalized AP, GWP,
and SCC rankings because of the low emissions of hydrogen production via thermochemical water splitting. Any enhancement on
energy and exergy efficiencies of this option in the long run would lower the hydrogen production costs, which would make
thermochemical water splitting have closer to ideal case rankings with highest possible efficiencies and lowest possible harmful
emissions and production costs.

3.1.6.3 Hydrogen Production From Photonic Energy


PV-based electrolysis, photocatalysis, and PEC cells are selected as photonic energy-based hydrogen production in this chapter. The
normalized ranking comparison of photonic energy powered hydrogen production options is presented in Fig. 10. Hydrogen
production from solar (photonic) energy is currently in its early research and development state. For that reason, among the
selected options from other primary energy resources for hydrogen production, photonic energy-based production options have
rankings for lower energy and exergy efficiencies and hydrogen production costs. On the other hand, photonic-based hydrogen
Hydrogen Production 25

Energy
efficiency
10

6 Exergy
AP
efficiency
4
Thermolysis
2
Thermochemical
0 water splitting
Biomass conversion

Biomass gasification

GWP Cost Biofuel reforming

Ideal

SCC
Fig. 9 Normalized ranking comparison of thermal-based hydrogen production methods. AP, acidification potential; GWP, global warming
potential; SSC, social cost of carbon.

Energy
efficiency
10

6 Exergy
AP efficiency
4

2 PV electolysis

0
Photocatalysis

Photoelectrochemical
GWP Cost method

Ideal

SCC
Fig. 10 Normalized ranking comparison of photonic-based hydrogen production methods. AP, acidification potential; GWP, global warming
potential; SSC, social cost of carbon.

production has very low CO2 and SO2 emissions and low SCC compared to the other selected options. As a result, environmental
and social impact evaluation of photonic energy-based hydrogen production options have nearly ideal case normalized rankings.
Amongst the selected photonic energy-based hydrogen production methods, PV electrolysis has the highest energy and exergy
efficiencies. Nevertheless, electrolysis also has the highest CO2 and SO2 emissions and SCC compared to the other photonic
energy-based options investigated in here. Photocatalysis has the lowest efficiencies and lowest hydrogen production cost. PEC
cells are in the earliest state of research and development, so they has very low efficiencies and high hydrogen production costs.
Nonetheless, PEC has nearly ideal case when emissions and social impact are considered although by far it has the highest
hydrogen production costs and significantly low efficiencies as can be seen from Fig. 10. Average normalized rankings show that
photocatalysis has the closest to ideal case performance with its highest average rankings. On the other hand, PV-based electrolysis
has the lowest normalized average rankings among selected photonic energy-based hydrogen production methods.
26 Hydrogen Production

3.1.6.4 Hydrogen Production From Hybrid Sources


Selected hybrid energy-based hydrogen production methods in this section are HTE, hybrid thermochemical cycles, biophotolysis,
photofermentation, artificial photosynthesis, and photoelectrolysis. These hybrid energy resources can be listed as electrothermal,
photobiochemical, and electrophotonic. The normalized ranking comparison of hybrid energy powered hydrogen production
options is presented in Fig. 11. On average, hybrid energy-based hydrogen production methods are more environmentally friendly
compared to hydrogen production powered by electrical, thermal, and photonic alone. As a result, hybrid energy-based hydrogen
production options have higher GWP, AP, and SCC rankings, and, therefore, are closer to the ideal case than the hydrogen
production options from single principal energy resources. In order to make hybrid energy-based hydrogen production methods
more viable both in technical and economic ways, their energy and exergy efficiencies should further be enhanced and their
hydrogen production costs should be reduced. Among the selected hybrid energy-based hydrogen production options, hybrid
thermochemical cycles have the highest energy and exergy efficiencies. On the other hand, artificial photosynthesis and photo-
electrolysis have the lowest energy and exergy efficiencies. Photofermentation is the cheapest hydrogen production method in this
category and hybrid thermochemical cycles have the highest production method. Photonic energy-based hybrid options (e.g.,
photofermentation, artificial photosynthesis, and photoelectrolysis) have the highest GWP, AP, and SCC rankings because they
have the lowest emissions among the selected options. The average normalized rankings of hybrid hydrogen production alter-
natives indicate that hybrid thermochemical cycles are closest to the ideal case with highest rankings. HTE has the lowest rankings
(least ideal option).

3.1.6.5 Overall Comparison


For overall comparison, average normalized rankings of GWP, AP, SCC, production cost, and energy and exergy efficiencies of each
method by their principal energy source are calculated. This is done with the intention of having an overall economic, technical,
social, and environmental evaluation of each principal energy resource. The results are provided in Table 10 and shown in Fig. 12.
The average normalized rankings indicate that hydrogen production from electrical energy has the highest energy efficiency and
is the cheapest option. On the other hand, hydrogen production methods powered by electrical energy, on average, have the

Energy
efficiency
10

6 Exergy
AP
efficiency
4
High temperature electrolysis
2
Hybrid thermochemical cycles
0
Biophotolysis
Photofermentation
Artificial photosynthesis
GWP Cost
Photoelectrolysis
Ideal

SCC
Fig. 11 Normalized ranking comparison of hybrid energy-based hydrogen production methods. AP, acidification potential; GWP, global warming
potential; SSC, social cost of carbon.

Table 10 Overall normalized ranking comparison of primary energy sources investigated in this study

Primary energy Energy Exergy Cost Social cost Global warming Acidification
source efficiency efficiency of carbon (SCC) potential (GWP) potential (AP)

Electrical 6.15 2.85 8.26 2.08 2.08 7.00


Thermal 5.04 4.06 7.69 7.08 7.08 3.94
Photonic 0.71 0.32 3.23 8.89 8.89 9.05
Hybrid 2.13 1.87 7.08 8.73 8.73 9.41
Ideal 10 10 10 10 10 10
Hydrogen Production 27

Energy efficiency
10

6
AP Exergy efficiency
4
Electrical
2
Thermal
0 Photonic

Hybrid

Ideal

GWP Cost

SCC
Fig. 12 Normalized ranking comparison of selected principal energy resources selected in this study. AP, acidification potential; GWP, global
warming potential; SSC, social cost of carbon.

highest GWP and SCC because of the high emissions amounts of plasma arc decomposition. Thermal energy powered hydrogen
production methods, on average, have the highest exergy efficiency and AP. This high AP is the direct result of the high SO2
emissions of biomass gasification. Photonic energy powered hydrogen production options, on average, are the most envir-
onmentally friendly ones, immediately followed by hybrid energy powered hydrogen production methods. Nevertheless, both
options have low efficiencies and high production costs. On average, hybrid hydrogen production methods have the highest
rankings (6.32/10), followed by thermal (5.82/10), photonic (5.18/10), and electrical (4.74/10) energy powered hydrogen
production.
Amongst the selected hydrogen production options, there is a wide ranging difference in technological advancement levels.
There are some methods already at technological maturity level. And some methods are at late research and development stages.
These methods have higher efficiencies and lower hydrogen production costs than the ones at early research and development
stages (for instance PEC cells). An additional significant issue is the readiness of large-scale production. For instance, fossil fuel
reforming and coal gasification can easily be used for large-scale hydrogen production, which significantly lowers the costs.
Clean, affordable, efficient, and reliable hydrogen production is the first step toward implementing sustainable hydrogen
energy systems. With the intention of reaching this target, there is a noteworthy volume of research supported by academia,
industry, and governments. The goal of these research activities is to enhance the technical and economic effectiveness of currently
available hydrogen production methods and identify novel promising technologies and materials for hydrogen production. The
production methods discussed in this chapter can either be employed alone, or integrated with other options to reach the targets of
sustainable hydrogen production. Due to their limited and nonrenewable reserves and resultant harmful gas emissions, fossil fuel-
based hydrogen production methods are not regarded as sustainable. Nevertheless, these options are considered for their help
during the transition to sustainable hydrogen production systems as novel renewable-based hydrogen production technologies are
getting more advanced.

3.1.7 Case Study: Hydrogen Production via Electrolysis

Electrolysis, which has been used since the late 1920s, is regarded as the first commercial method to generate pure hydrogen. In
electrolysis, external electricity is applied to split water into hydrogen and oxygen. In the 1960s, industrialized hydrogen pro-
duction had moved to fossil fuel powered options. To this date, fossil fuels are the main energy and material resources for
hydrogen production. Electrolysis-based hydrogen production currently makes up no more than 4% of the total global hydrogen
supply.
Many hydrogen production methods, including electrolysis, have been investigated heavily in the literature from economic,
environmental, technical, and social perspectives. Current global status of hydrogen energy systems have also been examined in
numerous literature studies, such as Momirlan and Veziroglu [50]. Ursua et al. [51] have comprehensively investigated some of
28 Hydrogen Production

state of the art of electrolysis technologies for sustainable hydrogen generation. Stojic et al. [52] have evaluated energy efficiencies
of different electrolysis technologies for hydrogen production.
The electrolyzer is the elementary component of a hydrogen production process via electrolysis. Electrolyzers can be coupled
together either in parallel or in series to establish an electrolyzer module. In most cases, the product gases of electrolysis (usually
hydrogen and oxygen) should be cooled down and compressed before storage for different end use applications [52]. In most
electrolysis processes, oxygen is not a desired product so it is not stored and most commonly is just emitted to the surrounding
areas. Also, water supply to the electrolyzer should be properly treated to eliminate impurities, and thus potential unwanted side
reactions.
Electrolyzers usually do not have moving parts, therefore, they do not require regular maintenance. In addition, electrolyzers
are quiet and modular; consequently they are good fits for distributed energy supply to residential, commercial, and industrial
units. Electrolyzers have been in use for different purposes for almost 100 years. However, significant technological and material
innovations are required to lower manufacturing, distribution, and installation costs of electrolyzers. In addition, electrolyzers’
system efficiencies should be enhanced significantly. Supporting efficient operation in moderate temperatures and pressures is also
required. In the literature, the three most commonly investigated electrolyzer types are alkaline, polymer electrolyte membrane,
and high-temperature SOE.

3.1.7.1 Alkaline Electrolyzers


There are two electrodes and a gas-tight diaphragm separating the anode and cathode compartments in alkaline electrolyzers. With
the introduction of the external direct electrical current, hydrogen is generated at the cathode and oxygen at the anode of the
electrolyzer. In alkaline electrolyzers, highly concentrated liquid electrolytes are used to enhance electrical conductivity. A common
liquid electrolyte used in alkaline electrolyzers is the aqueous solution of KOH at around 25 to 30 wt% concentration. NaOH or
NaCl solutions are less common electrolyte types used in alkaline electrolyzers. The major disadvantage of alkaline electrolyzers is
their corrosive environment.
In alkaline electrolyzers, hydrogen is produced at the cathode where water is reduced by generating hydroxide anions (OH).
The hydroxide anions cross the diaphragm from the cathode to the anode with the help of the electric field generated by the
external electricity supply. Then, in the anode, the hydroxide anions are recombined on the anode surface to generate oxygen. The
anode, cathode, and overall reactions of alkaline electrolyzers can be written as
Anode : 4OH -O2 þ 2H2 O þ 4e ð39Þ

Cathode : 4Hþ þ 4e -2H2 ð40Þ

Overall : 2H2 O-O2 þ 2H2 ð41Þ


The diaphragm used in alkaline electrolyzers to separate the anode and cathode also avoids the recombination of the product
gases. Henceforth, the diaphragm permits the ion transfer but is not permeable to gases. In earlier electrolyzers, about 3-mm-thick
asbestos was the selected diaphragm substance. Because of the operational limitations of asbestos, the electrolysis reaction cannot
take place at temperatures above 801C in conventional alkaline electrolyzers. In addition to this limitation, there are serious
environmental and health risks associated with asbestos use. Therefore novel materials are being investigated to substitute for
asbestos use in alkaline electrolyzers.
Electrodes are generally placed next to both sides of the diaphragm in alkaline electrolyzers. Nickel and nickel coated steel are
the most common anode materials and the cathode is usually activated steel. The steel cathode is activated via catalyst coating. The
space in between the anode and the cathode is around 5 mm in most electrolyzers [33].
Alkaline electrolyzers use mature technologies and they are reliable and safe. Therefore they are the most commonly used
electrolyzers in the world. The investment cost of alkaline electrolyzers is in between $1000 and $5000/kW and the actual cost
depends on the electrolyzer’s hydrogen production capacity. The operating costs of the alkaline electrolyzers have been decreasing
with recent technological advancements. This is mainly because of the decrease in energy input requirements due to increased
system efficiencies. The investment costs are also decreasing with reduced active surface area requirements as a result of enhanced
operating current densities. Another research topic for alkaline electrolyzers is the development of advanced membrane materials
to substitute asbestos. Inorganic ion exchange membranes are the most promising options mentioned in the literature [53,54]. In
addition, advanced alkaline electrolyzers operating at temperatures up to 1501C are currently being developed for scaling-up
hydrogen production via alkaline electrolyzers. In alkaline electrolyzers, the hydrogen and oxygen concentrations in corresponding
outlet streams are around 99.9 and 99.7 vol%, respectively, without using a supplementary gas treatment unit [55]. The product
gas concentrations can further be increased via catalytic conversion and adsorptive drying processes.

3.1.7.2 Polymer Electrolyte Membrane Electrolyzers


Polymer electrolyte membrane (PEM) electrolyzers are also known as proton exchange membrane or solid polymer electrolyte
(SPE) electrolyzers. Unlike alkaline electrolyzers, PEM electrolyzers do not need any liquid electrolytes. The gas-tight thin polymer
membrane acts as electrolyte in PEM electrolyzers. The most common membrane type used in PEM electrolyzers is less than 0.2-
mm-thick Nafion [51]. The lifetime of PEM electrolyzers depends on the lifetime of the PEM, which is shorter than that of alkaline
Hydrogen Production 29

electrolyzers. Combined together, the anode, cathode, and membrane are referred to as the membrane electrode assembly (MEA)
in PEM electrolyzers. The electrodes are generally made from noble metals such as platinum or iridium. The anode and cathode
reactions of PEM electrolyzers are:
1
Anode : H2 O- O2 þ 2Hþ þ 2e ð42Þ
2

Cathode : 2Hþ þ 2e -H2 ð43Þ


þ þ
Water oxidation reaction at the anode produces oxygen, electrons, and H ions. H ions cross the membrane to the cathode
side. Here, the reduction reaction of H þ ions closes the loop and produces hydrogen (Eqs. (34) and (35)). Commercial PEM
electrolyzers are available for low scale hydrogen production. The product purity of PEM electrolyzers is normally over 99.99 vol%
hydrogen. In some advanced PEM electrolyzers, this purity goes up to 99.999 vol% without the use of any supplementary
separation unit [51]. Moreover, low permeability of the polymeric membranes to any gas crossing minimizes any combustion risks
and fire hazards. A typical PEM electrolyzer has a hydrogen production capacity around 0.06 to 30 Nm3 H2/h, specific energy
demand around 6 to 8 kWh/Nm3 H2, and operates at temperatures and pressures up to 801C and 15 bar, respectively. Scaled up
PEM electrolyzers can have production capacities above 10 Nm3 H2/h with system efficiencies between 67 and 82% [51].
The polymeric membranes used in PEM electrolyzers respond quickly to inconsistent power supplies. For that reason, PEM
electrolyzers can operate with fluctuating external electricity supplies. Alkaline electrolyzers are different for that matter; the liquid
electrolytes used in alkaline electrolyzers are highly sensitive to power fluctuations [56]. There are certain disadvantages of PEM
electrolyzers, such as the high investment costs due to the use of expensive membranes and electrodes made from noble metals. In
addition, PEM electrolyzers should be scaled up in an efficient and cost effective manner to promote their wide commercial
applications [57].

3.1.7.3 Solid Oxide Electrolyzers


Steam electrolysis takes place in SOE at temperatures up to 10001C. Due to their high operating temperatures, SOEs have the
advantage of higher system efficiencies than alkaline or PEM electrolyzers [58]. In PEM electrolyzers, hydrogen is produced from
steam reduction at the cathode. The oxide anions (O2) are formed at the cathode and sent to the anode by passing the solid
electrolyte. At the anode, O2 ions are oxidized to oxygen gas and this reaction closes the electron loop. The most common
cathode material used in the literature is a cermet made from nickel and yttrium stabilized zirconia (YSZ); the solid electrolyte is
commonly YSZ and the anode is mostly perovskite [51].
Steam electrolysis is developed in order to lower the energy input requirements and operating costs of electrolyzers [59].
Theoretical calculations show that up to 40% of the energy requirement of electrolysis-based hydrogen production can be supplied
via the thermal energy stored in the steam when steam electrolysis is conducted at around 10001C. This could reduce the electricity
input requirement of the electrolyzer by about 25% [58]. For that reason, SOEs are regarded as more efficient compared to alkaline
and PEM electrolyzers. Therefore if a high-temperature heat source is available, SOEs are considered as promising hydrogen
production methods. The high-temperature heat source could be coming from anywhere such as nuclear reactors, geothermal,
solar, etc. SOEs are presently at research and development state and their demonstrations are conducted in laboratory scales with
hydrogen production capacities less than 5.7 Nm3/h and power ratings up to 18 kW.
The drawbacks of high-temperature electrolyzers mainly related to material selection when building the system. Fluctuations in
electricity supply cause heat losses and variations in operating temperatures, which could lead to microscale fractures in the
membrane. These fractures lower the lifetime of the overall system considerably. Therefore high-temperature electrolyzers are not
well suited to work with irregular energy sources such as renewable energies.

3.1.7.4 Environmental Impact Analysis of Electrolysis


In this section, environmental impacts of various electrolysis-based hydrogen production methods are comparatively assessed and
discussed. First, life cycle inventory results of wind-based electrolysis are presented to closely investigate each and every component
of an electrolysis process, which is the primary focus of this section. For this reason, well to wheel energy and material con-
sumption of wind-based electrolysis is considered. In well to wheel analysis, all resources needed to build and operate every
component of the wind-based electrolysis system are carefully evaluated. The main components are wind turbines, electrolyzers,
and hydrogen storage units.
In this section, a process with three 50 kW wind turbines and 30 Nm3/h hydrogen production capacity is taken as basis. The
electrical to hydrogen energy conversion efficiency of this electrolyzer is 85% based on the HHV of hydrogen. Before being sent to
end use points, the electrolyzer system compresses the product hydrogen up to pressures around 20 MPa.
Most important material and energy resources used in wind-based electrolysis are fossil fuels and some minerals as well as a
variety of metals. For example, iron is the most commonly used material resource in wind-based electrolysis systems, making up
37.4% of the overall material use in the entire lifetime of the overall system. Iron is used in building the wind turbines and
hydrogen storage units. Following iron, limestone use accounts for 35.5% of the total material resource use in wind turbines.
Limestone is used in the concrete bases of the wind turbines. Next, coal makes up 20.8% of the total material resource use, which is
employed predominantly to manufacture steel, iron, and concrete. Following iron, limestone, and coal, there are coal and natural
30 Hydrogen Production

Table 11 Accumulated lifetime material consumption in the selected wind-based electrolysis process

Material Total (g/kg H2) Wind turbines (%) Electrolysis (%) Storage (%)

Coal 214.7 68 5 27
Iron (Fe, ore) 212.2 64 6 30
Iron scrap 174.2 53 8 39
Limestone 366.6 96 1 3
Natural gas 16.2 72 15 13
Oil 48.3 76 13 11

Source: Reproduced from Bhandari R, Trudewind CA, Zapp P. Life cycle assessment of hydrogen production via electrolysis – a review. J Clean
Prod 2014;85:151–63.

Table 12 Accumulated lifetime emissions of the selected wind-based electrolysis process

Resource Total (g/kg H2) Wind turbines (%) Electrolysis (%) Storage (%)

CO2 950 78 4 18
CO 0.9 80 4 16
Methane 0.3 92 3 5
Nitrogen oxides 4.7 46 47 7
Nitride oxides 0.05 67 6 27
Nonmethane hydrocarbons 4.4 63 7 30
Particulates 28.7 94 1 5
Sulfur dioxide 6.1 62 26 12

Source: Reproduced from Bhandari R, Trudewind CA, Zapp P. Life cycle assessment of hydrogen production via electrolysis – a review. J Clean
Prod 2014;85:151–63.

gas, which account for 4.7 and 1.6% of the total material resources use in the entire lifetime of the system, respectively. Oil and gas
are principally employed in building the wind turbines. Water is consumed in both electrolyzer operation and various other
upstream processes. Approximately 45% of the water consumption in wind-based electrolysis takes place in the electrolyzer itself.
The rest is distributed as 38% during the manufacture of wind turbines and around 17% during the manufacture of hydrogen
storage units. Table 11 shows the accumulated material consumption in the selected wind-based electrolysis process.
The average energy consumption of the selected wind-based electrolysis system is 9.1 MJ/kg of hydrogen production. The most
significant portion (72.6%) of this energy consumption belongs to the manufacturing process of the wind turbines followed by the
hydrogen storage process (31.6%) including storage unit manufacturing, compression, etc. The actual electrolysis process itself
only accounts for 4.8% of the energy consumption in the overall lifetime of a wind-based electrolysis system.
The majority of the energy consumption, i.e., 72.6%, was in wind turbines manufacturing, whereas the share for electrolysis
and storage was 4.8 and 31.6%, respectively. CO2 emissions of this system are very high due to the emissions during the
manufacture and installation of wind turbines and hydrogen storage units. Table 12 summarized the overview of emissions for the
selected wind-based electrolysis process.

3.1.7.5 Global Warming Potential


GWP is the most commonly investigated environmental impact category in the literature. Fig. 13 presents the overview of the GWP
results of various electrolysis processes as they are provided in the literature [38]. It should be noted that these emissions results are
averages as the real emissions differ from region to region.
In terms of CO2 emissions, wind-based electrolysis seem to be the most environmentally benign one, which is followed by
hydro-based electrolysis. One reason of the emissions difference between the hydro- and wind-based electrolysis could be the
construction phase emissions of hydro power plants. Since hydro power plant constructions cause deforestation to build the dams,
channels, power plants, etc., their impact on the environment might be higher CO2 emissions compared to wind turbines with
similar capacities. In addition, over longer periods, biomass decomposition in large water dams might cause higher CO2 emis-
sions. Compared to wind and hydro-based electrolysis, solar PV and thermal as well as biomass-based electrolysis have higher CO2
emissions, and as a result, GWPs. PV panel manufacturing process is the main source of CO2 emissions of PV-based electrolysis.
Biomass gasification-based electrolysis might have a longer process chain, the net CO2 emissions of biomass-based processes are
usually quite low. However, GWP of biomass gasification-based electrolysis is still higher than solar thermal-based electrolysis.
One of the most significant concerns related to the environmental impact of electrolysis-based hydrogen production is about
how electricity is produced. In other words, clean hydrogen production via electrolysis requires clean electricity supply. For that
matter, renewable-based electricity is preferred over fossil fuel-based electricity. Fig. 13 shows that grid-based electrolysis has the
highest CO2 emissions since the grid electricity comes from fossil fuel combustion. The share of fossil fuels in current grid
Hydrogen Production 31

35

30

GWP (kg CO2 eq./kg H2)


25

20

15

10

0
Wind

thermal

Solar PV

Nuclear based

Hydroelectricity

With grid

Biomass
Solar

HTE
Fig. 13 Global warming potential (GWP) of different electrolysis-based hydrogen production methods. HTE, high-temperature electrolysis.

Hydrogen
Electrolyzer compression
production and and storage
operation 18%
4%

Wind turbine
production and
operation
78%

Fig. 14 Share of global warming potential (GWP) in wind electrolysis.

electricity supply mix is above 50% on average, which is quite high, resulting in high CO2 emissions. It should also be noted that
the variation in CO2 emissions data published in different literature studies are mainly because of the differences in system
boundaries (i.e., what is taken into account, what is not considered, etc.). In any case, Fig. 13 shows that when renewable energy
sources are used to produce electricity, electrolysis-based hydrogen production methods have very low CO2 emissions (GWP),
which makes them promising alternatives for sustainable hydrogen production.
Fig. 14 shows the life cycle analysis results of hydrogen production from wind-based electrolysis emissions. Here, it can be seen
that the CO2 emissions from electrolyzer manufacturing and operation is quite negligible (4%). On the other hand, the wind
turbine construction and operation is the main reason of GWP (78% of the overall CO2 emissions) and other impact categories
during the life cycle of wind turbine-based electrolysis. Hydrogen compression and storage also has significant contribution to the
GWP of wind-based electrolysis, accounting for 18% of the overall CO2 emissions in the entire lifetime.
In order to have more environmentally benign hydrogen production via electrolysis, renewable energy sources should be used
for electricity generation, as discussed earlier. Nuclear power-based thermochemical cycles also have low GWP results. Among the
available conventional technologies, coal gasification has the highest CO2 emissions. In order to make coal gasification-based
hydrogen production as environmentally benign as hydrogen from renewable energy-based electrolysis, CCS technologies must be
employed during coal mining and coal gasification. Following coal gasification, hydrogen production via SMR has very high CO2
emissions as well. In addition, there are significant energy losses in conversion from fossil fuel to electrical energy, then to
hydrogen energy, which lowers overall system efficiencies. Therefore thermal energy recovery from fossil fuel powered electricity
32 Hydrogen Production

generation units is a better energy source for hydrogen production, in terms of system efficiencies. As a result, it can be concluded
that fossil fuel powered thermal hydrogen production methods are more efficient with fewer CO2 emissions compared to
hydrogen production via fossil fuel powered electrolysis. Another interesting finding of Fig. 13 is that biomass gasification has
higher CO2 emissions than biomass-based electrolysis. The reason behind this would be where the electricity comes from, and
how the gasification emissions are handled.

3.1.7.6 Acidification Potential


AP, which is a measure of the SO2 emissions, is the second most heavily investigated environmental impact category in the
literature. Fig. 15 shows the average AP results of selected electrolysis-based hydrogen production methods, as stated in the
literature [31].
Similar to GWP results; wind, hydro, and nuclear-based electrolysis options produce hydrogen with fewer SO2 emissions. The
small variance in SO2 emissions among these methods is due to the emissions during the manufacture of system components of
these options. Solar PV-based electrolysis has relatively higher SO2 emissions compared to solar thermal-based hydrogen pro-
duction due to the high amounts of emissions during solar panel manufacturing process. Among the selected options presented in
Fig. 15, biomass-based hydrogen production has the highest SO2 emissions, even higher than the electrolysis with grid electricity,
which mainly comes from fossil fuel-based power sources. The main reason for high SO2 emissions of biomass-based hydrogen
production is the specific biomass processes such as fertilizers.
The amount of CO2 and SO2 emissions of electrolysis-based hydrogen production via nuclear energy is predominantly as a
result of the mining, construction, and installation processes during nuclear power plant operation. Some authors (such as Ref.
[60]) argue that these processes take relatively shorter times compared to the entire lifetime of nuclear power plants, therefore, the
negative impacts related to CO2 and SO2 emissions of the actual hydrogen production process can be negligible. Nevertheless, CCS
and acid gas neutralization methods are considered as the short-term solutions to mitigate CO2 and SO2 emissions. However, it
should also be remembered that other impact categories must be taken into account when selecting the most appropriate
hydrogen production technology. For example, nuclear waste management and its potential impacts should be addressed before
switching to nuclear-based electrolysis for hydrogen production.

3.1.8 System Integration in Hydrogen Production

Energy and cost efficient design and application of a reliable and environmentally friendly hydrogen energy system necessitates a
comprehensive and systematic approach to hydrogen production, distribution, delivery, and end use. There are complex rela-
tionships amongst the different system constituents that require cutting edge technologies to address whole system level challenges
and issues. There already exists a variety of challenges that strongly impact the production, distribution, delivery, storage, and end
use of hydrogen. In addition, there are challenges related to training for hydrogen energy systems and outreach to have public
acceptance of hydrogen energy systems:

35

30

25
AP (gSO2 eq./kg H2)

20

15

10

0
Wind

thermal

Solar PV

Nuclear based

Hydroelectricity

With grid

Biomass
Solar

HTE

Fig. 15 Acidification potential (AP) of selected electrolysis-based hydrogen production options. HTE, high-temperature electrolysis.
Hydrogen Production 33

• Need for commonly accepted and applied regional and global regulations, codes, and standards for hydrogen energy systems
including production, delivery, distribution, storage, and end use.
• Safety precautions.
• Consumer acceptance: providing the expected performance at a reasonable cost.
• Collaborative research and development.
• Technology validation through demonstrations by government, industry, and academia partnerships.
• Systems analyses to explore various pathways to widespread hydrogen energy use, including full cost accounting for all
competing energy systems.
• Ready accessibility to existing information on hydrogen technologies.
It should be noted that governments have important responsibilities as the early adopters of innovative and sustainable
hydrogen production, supply, distribution, and end use systems and as the developers of reliable, efficient, clean, and affordable
hydrogen energy infrastructure. System integration addresses ways in which different parts of a system work together from
technical, economic, and societal standpoints. In many cases, system optimization may require a distinctly different approach
from the optimization of a single part. Similarly, a systems focus makes it easier to identify key technical or market barriers in any
one part of the system that might impede the development of the whole. Optimization at the system level will require the
following:

• Coordination of technology development between hydrogen producers and end users who require hydrogen at a particular
purity and pressure.
• A strong, coordinated, and focused research and development program that includes breakthroughs in hydrogen storage,
production, and use that could influence how fast and in what way(s) a hydrogen economy develops.
• Efficient coordination of supply and demand to solve the perceived “chicken and egg” problem in transportation markets:
vehicle manufacturers wish to be assured of fuel supply, while suppliers wish to be assured of a market.
Finally, moving to widespread use of hydrogen as an energy carrier involves profound changes in how we view and use energy
as individuals and as a society. Actions in the following areas are essential to establishing the underlying set of “system level”
preconditions or the context for creating a hydrogen energy system:

• Government leadership to identify and sustain the required long-term activities.


• Adoption of policies that consistently incorporate the external costs of energy (such as energy supply security, air quality, and
global climate change) and provide a clear signal to industry and consumers.
• Development of domestic and international markets for hydrogen energy to harness the projected growth of energy demand in
developing nations over the next half century.

Appropriate rules, regulations, codes, and standards are significant enablers of the commercialization of any novel system or
product. Even though there are many end use applications for the extensive use of hydrogen as an industrial chemical, hydrogen
production and end use in energy systems are expected to necessitate totally different rules, regulations, codes, and standards.
There should be regional and state level committees focusing on the development of hydrogen codes and standards and effective
coordination to design, develop, and test innovative hydrogen production methods and end use technologies. These committees
should organize and coordinate various actions conducted by different organizations in order to develop and adopt codes required
by hydrogen production and end use technologies. These committees should communicate through the hydrogen communities
and work for the advancement of reliable and stable rules, regulations, codes, and standards to fast track the market introduction
and success of hydrogen production and end use systems.

3.1.9 Hydrogen Production Challenges

Sustainable hydrogen energy systems require reliable, affordable, clean, and efficient hydrogen production systems with secure and
abundant energy services. In order to accomplish the sustainable hydrogen vision, there are numerous challenges that must be
addressed first, such as:

• High production costs of hydrogen compared to other commonly used fuels: current hydrogen production is mostly based on
fossil fuel processing. However, the unit cost of energy stored in hydrogen is significantly higher than the unit cost of energy
stored in fossil fuels. Combined with the high storage, distribution, and refueling station costs, the overall price of hydrogen
paid by end users is a lot higher than that of any conventional fossil fuel. The reason why conventional fossil fuels are a lot
cheaper than hydrogen from any form of production, storage, and delivery is their large scale, very well developed, and
technologically mature production, distribution, storage, and end use infrastructures.
• Constraints on hydrogen infrastructure capacity due to low end user demands: even though there is a steady increase in
hydrogen demand by various industries, such as refineries and chemicals processing, the demand for hydrogen as an energy
carrier and fuel is currently at very low levels. There is a complex relationship between a well-developed hydrogen infrastructure
and high demand for hydrogen energy systems as they both depend on each other. A well-developed hydrogen infrastructure
covers all elements of hydrogen energy systems from production, storage, distribution, refueling, and end use technologies such
34 Hydrogen Production

as hydrogen fueled cars, internal combustion engines, turbines, etc. The challenge is that without a strong infrastructure, the
demand for hydrogen energy is low, and without a certain level of demand, the incentives to fully design, develop, build,
install, and operate a well-developed hydrogen infrastructure is quite low.
• High GHG emissions and lack of full competency of current hydrogen production methods: presently, most mature tech-
nologies for large-scale hydrogen production are based on fossil fuel processing and these options emit large quantities of
GHGs. In addition, these mature technologies and the novel hydrogen production options should be improved in such ways to
provide reliable, affordable, abundant, clean, and efficient hydrogen at high purities.
• Innovative hydrogen production options are at an early research and development stage and require substantial improvement:
most of these novel hydrogen production options are discussed in detail in this chapter. Innovative hydrogen production
options take advantage of renewable and abundant energy sources such as geothermal, hydro, solar, wind, nuclear, and
biomass. Examples of innovative hydrogen production technologies are thermochemical water splitting via heat recovery from
other processes (such as solar, nuclear, etc.), direct photon to hydrogen energy conversion via PEC cells, advanced bioengi-
neered processes mimicking photosynthesis, etc. These methods need long-term and continuous research and development
efforts to make hydrogen energy systems ready for the market.
• Design, development, implementation, and demonstration of innovative hydrogen energy systems are required for public
acceptance: the potential end users should be provided an elementary perception of all aspects of hydrogen energy systems,
such as different energy and material sources for hydrogen production and different methods to produce, store, distribute, and
use hydrogen. For that matter, product demonstrations are highly important to gain public acceptance of hydrogen energy
systems.

3.1.9.1 How Much Hydrogen Is Needed to Meet Our Energy Demand?


With the development and implementation of sustainable hydrogen energy systems from production to end use applications,
global hydrogen demand is expected to require much more hydrogen than the current amount of production. In order to meet the
energy demand of around 25 million homes or 100 million hydrogen fueled automobiles, 40 million tons of hydrogen is
projected to be needed annually. Therefore the following options are provided, each of which is capable of producing 40 million
tons of hydrogen per year.

3.1.9.1.1 Small scale distributed hydrogen production options

• Electrolysis: 1,000,000 small regional electrolysis units for district hydrogen production.
• Small reformers: 67,000 small reformers operating on fossil fuels for more centralized yet still district-based production.

3.1.9.1.2 Large-scale centralized hydrogen production options

• Coal or biomass gasification facilities: 140 plants similar to size of current fossil fuel operated power plants used to meet the
annual hydrogen demand.
• Thermochemical water splitting via nuclear energy: heat recovered from 100 nuclear power plants used to meet the annual
hydrogen demand.
• Fossil fuel powered refineries: 20 fossil fuel-based power plants, each of which have the capacity of a small scale refinery used to
meet the annual hydrogen demand.

3.1.9.1.3 Integrated hydrogen production approach


There are many critically important criteria that should be considered carefully when selecting the most appropriate hydrogen
production alternative for a given region or end use application. The hydrogen production methods discussed in this chapter are at
different stages of research and development and commercialization. A truly sustainable future with hydrogen energy systems will
probably require integrating current and future hydrogen production methods. An example of this integrated approach is the
incorporation of small and large scale or distributed and centralized hydrogen production. It should be noted that some of the
future hydrogen production methods might not even exist today. The following integrated hydrogen production alternatives are
only some possible sample scenarios to provide 40 million tons of hydrogen per year, which could definitely be improved by
novel approaches to hydrogen production:

• 100,000 Small scale distributed electrolyzers to generate up to 4 million tons of hydrogen.


• 15,000 Small reformers in refueling stations could generate up to 8 million tons of hydrogen.
• 30 Coal or biomass gasification plants could generate up to 8 million tons of hydrogen.
• Nuclear thermochemical water splitting plants could generate up to 4 million tons of hydrogen.
• 7 Large oil and gas SMR or gasification refineries could generate up to 16 million tons of hydrogen.
Hydrogen Production 35

3.1.10 Innovative Hydrogen Production Research at Clean Energy Research Laboratory

Comprehensive research is being conducted in novel H2 production systems at Clean Energy Research Laboratory (CERL) that can
lead to an environmentally benign and affordable H2 economy. Fig. 16 presents a diagram of the key H2 production research at
CERL which are (1) electrical, (2) thermal, and (3) light-based systems.

3.1.10.1 Electricity
Electrolysis is extensively utilized for large-scale H2 production. There are two types of electrolysis technologies: high- and low-
temperature. Low-temperature electrolysis only uses electricity to dissociate H2O. PV and geothermal electrolysis, and integrated
systems [61], are established for low-temperature H2 generation at UOIT. Various novel systems are designed and tested as well for
HTE that have reduced electrical input. Bicer et al. [62] have investigated the effect of solar spectra on PV efficiency. Hacatoglu et al.
[63] have assessed sustainability of a wind–H2 energy system assessed by comparing wind electrolysis to a conventional gas-fired
system. Soltani et al. [64] have conducted electrochemical analysis of an HCl/CuCl electrolyzer by considering equilibrium
thermodynamics. Bicer and Dincer [65] have developed a new solar and geothermal-based combined system for sustainable
electrolysis and H2 production.

3.1.10.2 Thermochemical
H2O dissociation is a promising way of H2 production via mid-high-temperature heat. A thermochemical H2O dissociation
process consists of a closed loop of heat driven reactions and all intermediate materials are recycled as H2O is split into H2 and O2.
These systems have the advantage of using waste and process heat, renewables, and nuclear heat as source of energy. Two main
thermochemical cycles (CuCl and MgCl) are being established at CERL at lab and pilot scales. The CuCl cycle is designed to utilize
nuclear process/waste heat aimed to support cogeneration in nuclear power plants. Ozcan and Dincer [66] have analyzed energy
and exergy efficiencies of a solar driven Mg–Cl hybrid thermochemical cycle for coproduction of power and hydrogen. Later,
Ozcan and Dincer [67] have comparatively assessed performances of different configurations of thermochemical MgCl cycle.
Ozbilen et al. [68] have developed and optimized a four step CuCl cycle for clean, affordable, and efficient H2 production. Ozcan
and Dincer [69] also have modeled a new four step MgCl cycle with dry HCl capture for more efficient H2 production.

3.1.10.3 Solar
Solar driven systems designed at CERL support photoelectrolysis, photocatalysis, and PEC. These are experimentally and con-
ceptually investigated in batch and continuous operation. Novel reactors are engineered and built for efficient, reliable, affordable,
and clean H2 production. Numerous integrated photonic-based systems are designed and investigated for large-scale production.
Several catalysts are evaluated for enhanced system performances.
In 2012, Zamfirescu et al. [70] have analyzed molecular charge transfer and quantum efficiency of a PEC hydrogen production
reactor. After that, Zamfirescu et al. [71] have investigated PEC chlorination of cuprous chloride with hydrochloric acid for
hydrogen production. In another study, Zamfirescu et al. [72] evaluated quantum efficiency modeling and system scaling-up
analysis of water splitting with Cd1xZnxS solid-solution photocatalyst. Zamfirescu and Dincer [73] have assessed a new integrated
solar energy system for hydrogen production.
An experimental study of hybrid photocatalytic H2O splitting reactor for an expanded range of the solar spectrum with CdS and
ZnS catalysts for H2 and O2 production is conducted at CERL [74]. Acar and Dincer [33] have reviewed and evaluated photo-
electrode coating materials and methods for PEC-based H2 production. Bicer and Dincer [75] have experimentally investigated a

Hydrogen production research


at UOIT

Electricity Thermochemical Solar light


• Low temperature electrolysis • Cu–CI cycle • Photoelectrolysis
• High temperature electrolysis • Mg–CI cycle • Photocatalysis
• Hybrid sulfur cycle • Photoelectrocatalysis

Hybrid thermochemical solar


• Hybrid photocatalytic Cu–CI hydrogen production

Fig. 16 Main areas of the H2 production research at Clean Energy Research Laboratory (CERL) at UOIT. Reproduced from Dincer I, Naterer GF.
Overview of hydrogen production research in the Clean Energy Research Laboratory (CERL) at UOIT. Int J Hydrog Energy 2014;39(35):20592–613.
36 Hydrogen Production

CERL novel
system
Photo-
electrochemical
Photo- (PEC)
catalysis(PC)

Photoelectrolysis (PE)

PV-electrolysis (PV-E)

Fig. 17 Evolution of innovation in hydrogen production driven by solar energy. CERL, Clean Energy Research Laboratory.

PV-coupled PEC-H2 production system at CERL. Acar and Dincer [76] have conducted experimental investigation of a hybrid PEC-
H2 production system. Evolution of photonic H2 production at CERL is presented in Fig. 17.
In addition, hybrid systems are integrated, designed, and tested at CERL to produce H2 in large scale and in a clean, efficient,
reliable, affordable manner. These hybrid systems are investigated by using energy, exergy, cost, and environment sides. An
innovative integrated solar system for large-scale (40 t/day) H2 production is reported [77].

3.1.11 Future Directions

Government, industry, and academic coordination on hydrogen production systems is required to lower overall costs, improve
efficiency, and reduce the cost of carbon sequestration. Better techniques are needed for both central station and distributed
hydrogen production. Efforts should focus on improving existing commercial processes such as SMR, multifuel gasification, and
electrolysis. Development should continue on advanced production techniques such as biological methods and clean nuclear or
solar powered direct photon to hydrogen energy-based water splitting.
The specific needs and actions required to address these barriers differ for each of the hydrogen production technologies. No
single technology meets all of the criteria of the ultimately reliable, affordable, clean, and efficient hydrogen production vision;
various combinations of the production technologies are likely to be used for different applications.

• Enact policies that foster both technology and market development: government support for research and development should
focus on developing advanced renewable and low carbon emitting methods plus carbon dioxide capture and sequestration
technologies.
• Improve gas separation and purification processes: the oxygen plant is one of the higher cost items in multifuel gasifiers;
lowering this cost will improve the economics of hydrogen production. Small, low cost, high efficiency hydrogen purification
methods are needed for distributed reformers that can generate hydrogen at residences or car refueling stations. Although some
purification technologies work well at large commercial sites, these are often difficult to scale down to the size needed for
distributed generation.
• Develop and demonstrate small reformers: small reformers that run on natural gas, propane, methanol, or diesel can provide
hydrogen to some of the first fleets and retail sales points, reducing overall costs. The technology also needs further refinement
for improved reliability, longer catalyst life, and integration with storage systems and fuel cells.
• Optimize and reduce costs of electrolyzers: efforts to improve the efficiency and lower the costs of electrolyzers must continue,
as this production method is ideal for distributed generation and could offer early market opportunities. Although electrolysis
is currently more expensive than thermal production, a better understanding of high-temperature and high-pressure electrolysis
could bring costs down. In distributed hydrogen systems, the hydrogen produced onsite often requires compression (to
pressures as high as 240 atm) for storage; high-pressure electrolysis could remove the need for this additional compression.

A near-term study should be conducted to develop measurable goals for novel hydrogen production options in terms of pro-
duction efficiency, emissions, operation and investment cost, and price. Specific goals will help to align and focus development efforts.

• Develop advanced renewable energy methods that do not emit carbon dioxide: photolytic processes use light energy to split
water and produce hydrogen, potentially offering lower costs and higher efficiencies for collecting solar energy. Semiconductors
that enable PEC splitting of water need to become more efficient and less susceptible to corrosion in water. Biological systems
may become a “low tech” way to provide hydrogen, but they are still in the early stages of development.
Hydrogen Production 37

• Develop advanced nuclear energy methods to produce hydrogen: research is needed to identify and develop methods for
economically producing hydrogen with nuclear energy, which would avoid carbon emissions. Thermochemical water splitting
using high-temperature heat from advanced nuclear reactors could be included in future nuclear plant designs.
• Develop methods for large-scale carbon dioxide capture and sequestration: a cost effective way to capture and sequester carbon
dioxide would facilitate the production of vast quantities of hydrogen with low carbon emissions. Capture systems would need
to be engineered into plant designs for SMRs and multifuel gasifiers to lower the overall systems costs.
• Demonstrate production technologies in tandem with applications: demonstrations are expensive, especially since there may
be little initial demand for the hydrogen produced. Demonstrations that integrate production technology with other elements
of the hydrogen infrastructure, including a market use, will be more cost effective. These demonstrations should highlight safety
and other benefits to stimulate market interest.

Demonstrations of hydrogen generation, purification, storage, dispensing, and fuel cell electricity generation should be pursued
in the short term in major metropolitan areas. For technologies that need larger scale testing and demonstration, an industrial scale
testing location should be developed to alleviate difficulties in finding acceptable sites.

3.1.12 Closing Remarks

This chapter comparatively evaluates and assesses environmental, financial, social, and technical performance of selected hydrogen
production methods. Electrical, thermal, photonic, electrothermal, photobiochemical, and electrophotonic are the primary energy
sources of these selected methods. Material resources of these methods are water, biomass, and fossil fuels. Six criteria are selected
for comparison purposes: GWP, AP, SCC, production cost, and energy and exergy efficiencies. The results of this study can be listed
as:

• Fossil fuel reforming has the highest (83%) and photocatalysis (less than 2%) has the lowest energy efficiency among selected
options. In general, photonic (solar)-based hydrogen production options have low energy efficiencies.
• Biomass gasification has the highest exergy efficiency (60%), followed by fossil fuel reforming (around 45–50%). Again,
photonic-based hydrogen production options have lowest exergy efficiencies compared to other selected options.
• The production cost evaluation shows that fossil fuel reforming ($0.75/kg H2), coal gasification ($0.92/kg H2), and plasma arc
decomposition ($0.85/kg H2) produce the cheapest hydrogen. On the other hand, as an early R&D phase method, PEC
hydrogen ($10.36/kg H2) is by far the most expensive one.
• GWP and AP of photonic-based hydrogen production methods are almost zero. As a result, these options have very low SCC.
On the other hand, fossil fuel reforming, plasma arc decomposition, biomass and coal gasification have very high GWP, AP,
and SCC among the selected options.
• The average normalized rankings of individual methods show that hybrid thermochemical cycles give clooset to sdtato#ideal
case results (7.57/10). This amount is the lowest for coal gasification (3.55/10).
• When selected methods are compared based on their primary energy sources, electrical-based hydrogen production show the
highest energy efficiency and lowest production cost. This method also gives the highest GWP and SCC.
• Thermal-based hydrogen production has the highest exergy efficiency and AP. Photonic-based hydrogen production gives the
lowest AP, GWP, and SCC.
• On average, hybrid hydrogen production methods have the highest rankings (6.32/10), followed by thermal (5.82/10),
photonic (5.18/10), and electrical (4.74/10)-based hydrogen production.

Research, development, and demonstrations are needed to improve and expand methods of economically producing hydrogen.
Production costs need to be lowered, efficiency improved, and carbon sequestration techniques developed. Better techniques are
needed for both central station and distributed hydrogen production. Efforts should focus on existing commercial processes such
as SMR, multifuel gasifiers, and electrolyzers, and on the development of advanced techniques such as biomass pyrolysis and
nuclear thermochemical water splitting, PEC electrolysis, and biological methods.

References

[1] International Energy Agency. 2016 Key world energy statistics: technical report. Available from: https://www.iea.org/publications/freepublications/publication/KeyWorld2016.
pdf; 2016 [accessed 10.02.16].
[2] Acar C, Dincer I. Comparative environmental impact evaluation of hydrogen production methods from renewable and nonrenewable sources. In: Dincer I, Colpan CO,
Kadioglu F, editors. Causes, impacts and solutions to global warming. New York, NY: Springer; 2013. p. 493–514.
[3] Dincer I. Environmental and sustainability aspects of hydrogen and fuel cell systems. Int J Energy Res 2007;31(1):29–55.
[4] Ryland DK, Li H, Sadhankar RR. Electrolytic hydrogen generation using CANDU nuclear reactors. Int J Energy Res 2007;31(12):1142–55.
[5] Dincer I, Balta MT. Potential thermochemical and hybrid cycles for nuclear‐based hydrogen production. Int J Energy Res 2011;35(2):123–37.
[6] Muradov NZ, Veziroǧlu TN. From hydrocarbon to hydrogen–carbon to hydrogen economy. Int J Hydrog Energy 2005;30(3):225–37.
[7] Levin DB, Chahine R. Challenges for renewable hydrogen production from biomass. Int J Hydrog Energy 2010;35(10):4962–9.
[8] Awad AH, Veziroǧlu TN. Hydrogen versus synthetic fossil fuels. Int J Hydrog Energy 1984;9(5):355–66.
[9] Yilanci A, Dincer I, Ozturk HK. A review on solar-hydrogen/fuel cell hybrid energy systems for stationary applications. Prog Energy Combust Sci 2009;35(3):231–44.
38 Hydrogen Production

[10] Lodhi MAK. Hydrogen production from renewable sources of energy. Int J Hydrog Energy 1987;12(7):461–8.
[11] Lodhi MAK. Helio-hydro and helio-thermal production of hydrogen. Int J Hydrog Energy 2004;29(11):1099–113.
[12] Miltner A, Wukovits W, Pröll T, Friedl A. Renewable hydrogen production: a technical evaluation based on process simulation. J Clean Prod 2010;18:S51–62.
[13] Lemus RG, Duart JMM. Updated hydrogen production costs and parities for conventional and renewable technologies. Int J Hydrog Energy 2010;35(9):3929–36.
[14] Alstrum-Acevedo JH, Brennaman MK, Meyer TJ. Chemical approaches to artificial photosynthesis. 2. Inorg Chem 2005;44(20):6802–27.
[15] Tanksale A, Beltramini JN, Lu GM. A review of catalytic hydrogen production processes from biomass. Renew Sustain Energy Rev 2010;14(1):166–82.
[16] Karunadasa HI, Chang CJ, Long JR. A molecular molybdenum-oxo catalyst for generating hydrogen from water. Nature 2010;464(7293):1329–33.
[17] El-Bassuoni AM, Sheffield JW, Veziroglu TN. Hydrogen and fresh water production from sea water. Int J Hydrog Energy 1982;7(12):919–23.
[18] Ni M, Leung MK, Sumathy K, Leung DY. Potential of renewable hydrogen production for energy supply in Hong Kong. Int J Hydrog Energy 2006;31(10):1401–12.
[19] Fulcheri L, Probst N, Flamant G, et al. Plasma processing: a step towards the production of new grades of carbon black. Carbon 2002;40(2):169–76.
[20] Gaudernack B, Lynum S. Hydrogen from natural gas without release of CO2 to the atmosphere. Int J Hydrog Energy 1998;23(12):1087–93.
[21] Baykara SZ. Experimental solar water thermolysis. Int J Hydrog Energy 2004;29(14):1459–69.
[22] Balta MT, Dincer I, Hepbasli A. Thermodynamic assessment of geothermal energy use in hydrogen production. Int J Hydrog Energy 2009;34(7):2925–39.
[23] Rand DAJ, Dell RM. Fuels – hydrogen production| coal gasification. Encycl Electrochem Power Sources 2009;29:276–92.
[24] Acar C, Dincer I, Zamfirescu C. A review on selected heterogeneous photocatalysts for hydrogen production. Int J Energy Res 2014;38(15):1903–20.
[25] Acar C, Dincer I, Naterer GF. Review of photocatalytic water‐splitting methods for sustainable hydrogen production. Int J Energy Res 2016;40(11):1449–73.
[26] Quan X, Yang S, Ruan X, Zhao H. Preparation of titania nanotubes and their environmental applications as electrode. Environ Sci Technol 2005;39(10):3770–5.
[27] Rabbani M, Dincer I, Naterer GF. Efficiency assessment of a photo electrochemical chloralkali process for hydrogen and sodium hydroxide production. Int J Hydrog
Energy 2014;39(5):1941–56.
[28] Acar C, Dincer I. Analysis and assessment of a continuous-type hybrid photoelectrochemical system for hydrogen production. Int J Hydrog Energy 2014;39(28):15362–72.
[29] Koutrouli EC, Kalfas H, Gavala HN, et al. Hydrogen and methane production through two-stage mesophilic anaerobic digestion of olive pulp. Bioresource Technol
2009;100(15):3718–23.
[30] Das D, Veziroglu TN. Advances in biological hydrogen production processes. Int J Hydrog Energy 2008;33(21):6046–57.
[31] Holladay JD, Hu J, King DL, Wang Y. An overview of hydrogen production technologies. Catal Today 2009;139(4):244–60.
[32] Kotay SM, Das D. Biohydrogen as a renewable energy resource – prospects and potentials. Int J Hydrog Energy 2008;33(1):258–63.
[33] Acar C, Dincer I. A review and evaluation of photoelectrode coating materials and methods for photoelectrochemical hydrogen production. Int J Hydrog Energy 2016;41
(19):7950–9.
[34] Köne AÇ, Büke T. Forecasting of CO2 emissions from fuel combustion using trend analysis. Renew Sustain Energy Rev 2010;14(9):2906–15.
[35] Abánades A. The challenge of hydrogen production for the transition to a CO2-free economy. Agron Res 2012;10(1):11–6.
[36] Guinée JB, Gorrée M, Heijungs R, et al. Life cycle assessment: an operational guide to the ISO standards; part 3: scientific background. Leiden: Centre for Environmental
Studies (CML), Leiden University; 2001.
[37] Ozbilen A, Dincer I, Rosen MA. Life cycle assessment of hydrogen production via thermochemical water splitting using multi-step Cu–Cl cycles. J Clean Prod
2012;33:202–16.
[38] Bhandari R, Trudewind CA, Zapp P. Life cycle assessment of hydrogen production via electrolysis – a review. J Clean Prod 2014;85:151–63.
[39] Ozbilen A, Dincer I, Rosen MA. Comparative environmental impact and efficiency assessment of selected hydrogen production methods. Environ Impact Assess Rev
2013;42:1–9.
[40] Kopp RE, Mignone BK. The U.S. government’s social cost of carbon estimates after their first two years: pathways for improvement. Economics 2012;6:1–43.
[41] Parthasarathy P, Narayanan KS. Hydrogen production from steam gasification of biomass: influence of process parameters on hydrogen yield–A review. Renew Energy
2014;66:570–9.
[42] Uddin MN, Daud WW, Abbas HF. Potential hydrogen and non-condensable gases production from biomass pyrolysis: insights into the process variables. Renew Sustain
Energy Rev 2013;27:204–24.
[43] Ngoh SK, Njomo D. An overview of hydrogen gas production from solar energy. Renew Sustain Energy Rev 2012;16(9):6782–92.
[44] Trainham JA, Newman J, Bonino CA, Hoertz PG, Akunuri N. Whither solar fuels? Curr Opin Chem Eng 2012;1(3):204–10.
[45] Ismail AA, Bahnemann DW. Photochemical splitting of water for hydrogen production by photocatalysis: a review. Sol Energy Mater Sol Cells 2014;128:85–101.
[46] Singh L, Wahid ZA. Methods for enhancing bio-hydrogen production from biological process: a review. J Ind Eng Chem 2015;21:70–80.
[47] Ibrahim N, Kamarudin SK, Minggu LJ. Biofuel from biomass via photo-electrochemical reactions: an overview. J Power Sources 2014;259:33–42.
[48] Bičáková O, Straka P. Production of hydrogen from renewable resources and its effectiveness. Int J Hydrog Energy 2012;37(16):11563–78.
[49] Dincer I, Zamfirescu C. Sustainable hydrogen production options and the role of IAHE. Int J Hydrog Energy 2012;37(21):16266–86.
[50] Momirlan M, Veziroglu TN. Current status of hydrogen energy. Renew Sustain Energy Rev 2002;6(1):141–79.
[51] Ursua A, Gandia LM, Sanchis P. Hydrogen production from water electrolysis: current status and future trends. Proc IEEE 2012;100(2):410–26.
[52] Stojic ́ DL, Marčeta MP, Sovilj SP, Miljanic ́ ŠS. Hydrogen generation from water electrolysis – possibilities of energy saving. J Power Sources 2003;118(1):315–9.
[53] Vandenborre H, Leysen R, Baetslé LH. Alkaline inorganic-membrane-electrolyte (IME) water electrolysis. Int J Hydrog Energy 1980;5(2):165–71.
[54] Vermeiren P, Adriansens W, Moreels JP, Leysen R. Evaluation of the Zirfons separator for use in alkaline water electrolysis and Ni-H2 batteries. Int J Hydrog Energy
1998;23(5):321–4.
[55] Turner JA. Sustainable hydrogen production. Science 2004;305(5686):972–4.
[56] Edwards PP, Kuznetsov VL, David WI, Brandon NP. Hydrogen and fuel cells: towards a sustainable energy future. Energy Policy 2008;36(12):4356–62.
[57] Grigoriev SA, Porembsky VI, Fateev VN. Pure hydrogen production by PEM electrolysis for hydrogen energy. Int J Hydrog Energy 2006;31(2):171–5.
[58] Brisse A, Schefold J, Zahid M. High temperature water electrolysis in solid oxide cells. Int J Hydrog Energy 2008;33(20):5375–82.
[59] Salzano FJ, Skaperdas G, Mezzina A. Water vapor electrolysis at high temperature: systems considerations and benefits. Int J Hydrog Energy 1985;10(12):801–9.
[60] Utgikar V, Thiesen T. Life cycle assessment of high temperature electrolysis for hydrogen production via nuclear energy. Int J Hydrog Energy 2006;31(7):939–44.
[61] Ratlamwala TAH, Dincer I. Comparative efficiency assessment of novel multi-flash integrated geothermal systems for power and hydrogen production. Appl Therm Eng
2012;48:359–66.
[62] Bicer Y, Dincer I, Zamfirescu C. Effects of various solar spectra on photovoltaic cell efficiency and photonic hydrogen production. Int J Hydrog Energy 2016;41
(19):7935–49.
[63] Hacatoglu K, Dincer I, Rosen MA. Sustainability of a wind-hydrogen energy system: assessment using a novel index and comparison to a conventional gas-fired system.
Int J Hydrog Energy 2016;41(19):8376–85.
[64] Soltani R, Dincer I, Rosen MA. Electrochemical analysis of an HCl (aq)/CuCl (aq) electrolyzer: equilibrium thermodynamics. Int J Hydrog Energy 2016;41(19):7835–47.
[65] Bicer Y, Dincer I. Development of a new solar and geothermal based combined system for hydrogen production. Sol Energy 2016;127:269–84.
[66] Ozcan H, Dincer I. Energy and exergy analyses of a solar driven Mg–Cl hybrid thermochemical cycle for co-production of power and hydrogen. Int J Hydrog Energy
2014;39(28):15330–41.
[67] Ozcan H, Dincer I. Comparative performance assessment of three configurations of magnesium–chlorine cycle. Int J Hydrog Energy 2015;41(2):845–56.
[68] Ozbilen A, Dincer I, Rosen MA. Development of a four-step Cu–Cl cycle for hydrogen production – part II: multi-objective optimization. Int J Hydrog Energy 2016;41
(19):7826–34.
Hydrogen Production 39

[69] Ozcan H, Dincer I. Modeling of a new four-step magnesium–chlorine cycle with dry HCl capture for more efficient hydrogen production. Int J Hydrog Energy 2016;41
(19):7792–801.
[70] Zamfirescu C, Dincer I, Naterer GF. Molecular charge transfer and quantum efficiency analyses of a photochemical reactor for hydrogen production. Int J Hydrog Energy
2012;37(12):9537–49.
[71] Zamfirescu C, Naterer GF, Dincer I. Photo-electro-chemical chlorination of cuprous chloride with hydrochloric acid for hydrogen production. Int J Hydrog Energy 2012;37
(12):9529–36.
[72] Zamfirescu C, Dincer I, Naterer GF, Banica R. Quantum efficiency modeling and system scaling-up analysis of water splitting with Cd1xZnxS solid-solution photocatalyst.
Chem Eng Sci 2013;97:235–55.
[73] Zamfirescu C, Dincer I. Assessment of a new integrated solar energy system for hydrogen production. Sol Energy 2014;107:700–13.
[74] Baniasadi E, Dincer I, Naterer GF. Hybrid photocatalytic water splitting for an expanded range of the solar spectrum with cadmium sulfide and zinc sulfide catalysts. Appl
Catal A: Gen 2013;455(1):25–31.
[75] Bicer Y, Dincer I. Experimental investigation of a PV-Coupled photoelectrochemical hydrogen production system. Int J Hydrog Energy 2017;42(4):2512–21.
[76] Acar C, Dincer I. Experimental investigation and analysis of a hybrid photoelectrochemical hydrogen production system. Int J Hydrog Energy 2017;42(4):2504–11.
[77] UOIT/Phoenix Canada Oil Company Limited. NSERC-CRD project: progress report; 2013.

Futher Reading

Dincer I, Joshi AS. Solar based hydrogen production systems. Springer; 2013.
Dincer I, Midilli A, Kucuk H, editors. Progress in exergy, energy, and the environment. Springer; 2014.
Dincer I, Rosen MA. Exergy: energy, environment and sustainable development. Newnes; 2012.
Dincer I, Zamfirescu C. Sustainable hydrogen production. Elsevier; 2016.
Gandia LM, Arzamedi G, Diéguez PM, editors. Renewable hydrogen technologies: production, purification, storage, applications and safety Newnes; 2013.
Gangloff RP, Somerday BP, editors. Gaseous hydrogen embrittlement of materials in energy technologies: mechanisms, modelling and future developments. Elsevier; 2012.
Granqvist CG, editor. Materials science for solar energy conversion systems. vol. 1. Elsevier; 2013.
Gratzel M, editor. Energy resources through photochemistry and catalysis. Elsevier; 2012.
Hoffmann P, Dorgan B. Tomorrow's energy: hydrogen, fuel cells, and the prospects for a cleaner planet. MIT Press; 2012.
Hoffmann P. A history of hydrogen energy: a BIT of tomorrow's energy. MIT Press; 2014.
Justi EW. A solar – hydrogen energy system. Springer Science & Business Media; 2012.
Ohta T, editor. Solar-hydrogen energy systems: an authoritative review of water-splitting systems by solar beam and solar heat: hydrogen production, storage and utilisation
Elsevier; 2013.
Sherif SA, Goswami DY, Stefanakos EL, Steinfeld A, editors. Handbook of hydrogen energy CRC Press; 2014.
Veziroglu T, editor. Hydrogen energy. Springer Science & Business Media; 2012.
Williams LO. Hydrogen power: an introduction to hydrogen energy and its applications. Elsevier; 2013.
Winter CJ, Nitsch J, editors. Hydrogen as an energy carrier: technologies, systems, economy. Dordrecht: Springer Science & Business Media; 2012.
Zini G, Tartarini P. Solar hydrogen energy systems: science and technology for the hydrogen economy. Springer Science & Business Media; 2012.

Relevant Websites

https://www.airliquide.com/science-new-energies/hydrogen-energy
Air Liquide.
http://www.airproducts.com/industries/Energy/Hydrogen-Energy.aspx
Air Products.
http://www.alternative-energy-news.info/technology/hydrogen-fuel/
Alternative Energy News.
http://www.alternative-energy-tutorials.com/energy-articles/hydrogen-energy.html
Alternative Energy Tutorials.
http://www.afdc.energy.gov/fuels/hydrogen.html
Alternative Fuels Data Center.
http://bahcesehir.edu.tr/
Bahçeşehir University.
http://www.bahcesehir.edu.tr/icerik/3127-energy-systems-engineering
Bahçeşehir University – Energy Systems Engineering.
http://www.azocleantech.com/article.aspx?ArticleID=29
Clean Tech.
https://climatechangeconnection.org/solutions/alternate-energy-sources/hydrogen-energy/
Climate Change Connection.
http://www.conserve-energy-future.com/hydrogenenergy.php
Conserve Energy Future.
https://www.eia.gov/energyexplained/index.cfm?page=hydrogen_use
Energy Information Administration.
http://energystorage.org/energy-storage/technologies/hydrogen-energy-storage
Energy Storage Association.
http://www.fchea.org/hydrogen/
Fuel Cell and Hydrogen Energy Association.
http://www.gcgw.org/
Global Conference on Global Warming.
http://hydrogenenergycalifornia.com/
Hydrogen Energy California.
40 Hydrogen Production

http://www.hydrogenenergycenter.org/
Hydrogen Energy Center.
http://heshydrogen.com/
Hydrogen Energy Systems, LLC.
http://hydrogeneurope.eu/
Hydrogen Europe.
http://www.iahe.org/
International Association for Hydrogen Energy.
http://www.ich2p-2017.org/
International Conference on Hydrogen Production.
http://www.ieees9.fesb.unist.hr/
International Exergy, Energy, and Environment Symposium.
https://www.joiscientific.com/
JOI Scientific.
http://www.lindeus.com/en/innovations/hydrogen_energy/hydrogen_energy_applications/index.html
Linde Group.
http://www.the-linde-group.com/en/clean_technology/clean_technology_portfolio/hydrogen_energy_h2/index.html
Linde Group.
http://www.mcphy.com/en/markets/hydrogen-energy/
McPhy.
http://www.merlin.unsw.edu.au/energyh/about-hydrogen-energy/
Merlin.
https://www.nrel.gov/workingwithus/eds-hydrogen.html
National Renewable Energy Laboratory.
http://hidrojenteknolojileri.org/
2nd International Hydrogen Technologies Congress.
http://www.renewableenergyworld.com/hydrogen/tech.html
Renewable Energy World.
https://www.industry.siemens.com/topics/global/en/pem-electrolyzer/silyzer/hydrogen-green-energy-of-the-future/pages/hydrogen-the-green-energy-of-the-future.aspx
Siemens.
https://www.studentenergy.org/topics/hydrogen
Student Energy.
http://www.iae.or.jp/e/research-groups/research-and-development-division/hydrogen-energy-group/
The Institute of Applied Energy.
https://www.uoit.ca/
University of Ontario Institute of Technology.
http://cerl.uoit.ca/
University of Ontario Institute of Technology – Clean Energy Research Laboratory.

You might also like