Journal of Cleaner Production: Rami S. El-Emam, Hasan Ozcan, Calin Zam Firescu

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Journal of Cleaner Production 262 (2020) 121424

Contents lists available at ScienceDirect

Journal of Cleaner Production


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

Updates on promising thermochemical cycles for clean hydrogen


production using nuclear energy
Rami S. El-Emam a, *, Hasan Ozcan b, Calin Zamfirescu c
a
Faculty of Engineering, Mansoura University, Mansoura, Egypt
b
Faculty of Engineering, Ankara Yildirim Beyazit University, Ankara, Turkey
c
School of Science and Engineering Technology, Durham College, Oshawa, Ontario, Canada

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

Article history: Hydrogen is having a worldwide momentum as a clean versatile energy solution and this is the time for
Received 17 December 2019 nuclear hydrogen production technologies to play its role in securing a clean and affordable energy
Received in revised form future. For over four decades, several nations have been investigating the potential of tens of thermo-
23 March 2020
chemical cycles for efficient and sustainable hydrogen production. These cycles require high quality heat
Accepted 29 March 2020
Available online 2 April 2020
(i.e. heat at high temperatures) solely or along with electric power (i.e. hybrid thermochemical cycles)
both of which can be provided by nuclear energy. This paper delivers a highlight on the potential of
^ as de
Handling Editor: Cecilia Maria Villas Bo nuclear energy for hydrogen production. It also discusses the main features of five of the promising
Almeida thermochemical cycles that are considered for integration with nuclear power plants. Furthermore, the
paper highlights the current status and advances of R&D in these thermochemical cycles as well as cost
Keywords: estimation and the main safety consideration for large-scale nuclear hydrogen production using different
Hydrogen nuclear-driver technologies.
Nuclear © 2020 Elsevier Ltd. All rights reserved.
Economics
Cogeneration
Safety

1. Introduction decision makers and stakeholders towards nuclear hydrogen


production.
In June 2019, the International Energy Agency (IEA) released a During the United Nations Climate Change Conference (COP 25)
landmark report titled the future of hydrogen: seizing today’s op- held in Madrid, Spain in December 2019, the International Atomic
portunities (IEA, 2019a). In the forward of the report executive Energy Agency (IAEA) Director General Rafael M. Grossi stated that
summary, Dr. Fatih Birol, Executive Director at the IEA, stated that greater use of low-carbon nuclear power is needed to ensure the
“hydrogen is today enjoying unprecedented momentum. The world global transition to clean energy. He also mentioned that this view
should not miss this unique chance to make hydrogen an important is shared by many of the IAEA’s 171 member states. During the last
part of our clean and secure energy future”. The report provides quarter of 2019, there was the announcement of three projects at
guidance on the future development of hydrogen production and the USA by three different utility companies to demonstrate
brings key recommendations for hydrogen production scale up to hydrogen production using energy of three nuclear power plants
stakeholders. This report was preceded by another report from the during 2020 and 2021. These projects are funded by the US
IEA on the role of nuclear power in a clean energy system (IEA, Department of Energy (DoE) with collaboration of several national
2019b). Since the beginning of the nuclear energy era, it was al- laboratories. Another project was announced in UK around the
ways considered as clean and sustainable energy source for inte- same time.
gration with hydrogen technologies for large-scale hydrogen Most of currently operating nuclear power plants are of Pres-
production. Currently, there is a growing interest supported by surized Water Reactor (PWR) technology where the maximum
temperature does not exceed 350  C (IAEA, 2019). Incorporating
hydrogen production using low-temperature hydrogen technology
(i.e. conventional electrolysis) to operating nuclear power plants of
* Corresponding author. water-cooled reactor type will improve the competitiveness of
E-mail addresses: rami.elemam@gmail.com (R.S. El-Emam), hozcan@ybu.edu.tr
nuclear industry through the cogeneration of hydrogen along with
(H. Ozcan), calin.zamfirescu@durhamcollege.ca (C. Zamfirescu).

https://doi.org/10.1016/j.jclepro.2020.121424
0959-6526/© 2020 Elsevier Ltd. All rights reserved.
2 R.S. El-Emam et al. / Journal of Cleaner Production 262 (2020) 121424

electricity. In addition, the use of hydrogen production for energy technology. Since cycle operating temperature and electrical en-
storage enables the flexible and load-following operation of these ergy consumption are two main indicators in the selection of
reactors. This also facilitates the feasible integration of nuclear and thermochemical cycles, some selected thermochemical cycles are
renewable energies in sustainable energy systems. comparatively presented in Fig. 2 with respect to these two in-
Many advanced reactors are proposed to increase thermal effi- dicators. However, this figure shows that comparing these cycles
ciency of nuclear power plants by increasing the maximum reactor based on their maximum temperature values and electrical energy
temperature. Such reactors are feasible to provide the required consumption may not be solely used to determine the most feasible
source of energy for driving thermochemical hydrogen production option for hydrogen production. One can see the pure thermo-
cycles. Fig. 1 gives an idea on the possibilities of nuclear integration chemical NaeOeH cycle in which the rate of reaction may not be
with hydrogen technologies based on the ranges of operating feasible at the mentioned maximum temperature. Another pure
temperature of the different promising hydrogen technologies and thermochemical cycle is the U-Eu-Br cycle whose chemical content
the quality of the heat provided by the different reactors. has radioactive components and may bring higher safety and se-
There are several publications in the literature tackling the curity concerns. Kinetics study is also inevitable for every individ-
development of specific elements in each of the hydrogen ther- ual reaction in this cycle. The Zn/ZnO cycle has been successfully
mochemical cycles discussed in the current paper. However, there
is lack of publications gathering the status of progress on these
cycles. This paper highlights the main features and recent advances
of some of the promising thermochemical cycles that possess po-
tential of integration with nuclear power plants for large-scale
hydrogen production. It also discusses the economics of consid-
ering such technologies when available for large-scale production.

2. Thermochemical cycles for hydrogen production

The main motivation behind development of thermochemical


cycles is to reduce the high cost associated with direct electrolysis
of water which is an electric-energy intensive process. Pure ther-
mochemical cycles require heat input at high temperatures for
water splitting process. Thermochemical cycles are seen to reduce
hydrogen generation cost and operate at higher efficiency. How-
ever, high temperature requirements jeopardize the safety of the
process and bring new challenges to the scene such as material
compatibility and operation durations. Therefore, hybrid cycles
have been proposed where electrical energy is utilized along with
thermal energy. This leads to a reduction in the high temperature
requirements and elimination of the associated challenges. In this
paper, the term thermochemical cycle is sometimes used to refer to Fig. 2. Maximum temperatures and theoretical electrical energy requirements of
pure and/or hybrid thermochemical cycles as a general category of selected hydrogen production methods.

Fig. 1. Potential of nuclear reactor technologies for integration with hydrogen production technologies based on ranges of operating temperature.
R.S. El-Emam et al. / Journal of Cleaner Production 262 (2020) 121424 3

demonstrated at high temperature using concentrated solar energy Laboratories and General Atomics from the United States and the
(Steinfeld, 2002) and it requires very high temperature for com- Atomic Energy Commission (CEA) of France to construct a labora-
plete operation. Fig. 2 also shows the two Sulfur-based cycles. The tory scale process for the production of 100e200 L of hydrogen per
pure thermochemical SeI cycle has some chemistry complexity and hour.
material compatibilities issues. The chemistry complexity is The cycle has been considered by the Japan Atomic Energy
reduced in the HyS cycle at the cost of electrical energy con- Agency (JAEA) for scale-up. Starting with a proof of operation
sumption. Both cycles have the same range of operating tempera- concept in 1997 by 24 h of operation with 1 L of hydrogen pro-
tures. The hybrid thermochemical cycles: HyS, CueCl, CaeBr and duction per hour at a laboratory scale test of the process (Nakajima
MgeCl lie on either side of the diagonal line in Fig. 2, where they et al., 1998), and through several solid achievements including one-
may be noted to be more realistic options. One may comment on week continuous production of 30 L hydrogen per hour using a
this representation as follows: cycles at the lower side of the di- bench-scale glass-based test system in 2004 (Kubo et al., 2004),
agonal line may possess chemistry and reaction rate problems until the construction and operation of a test facility at JAEA inte-
while the ones at the upper side may possess economic infeasi- grating the three sections of the cycle: HI, Bunsen, and H2SO4
bility. Some recent advances in selected thermochemical cycles are sections; and successfully operating it for 8 h with 10 L/h of
discussed in next subsections. hydrogen production in February 2016 (Noguchi et al., 2017), fol-
lowed by successful 31 h of operation with hydrogen production
rate of 20 L/h later in October of the same year. After resolving the
2.1. SeI cycles
issue of solid precipitation from I2 in the HI decomposition section,
another milestone was achieved, as announced in the JAEA press
SeI cycle is the most developed pure thermochemical hydrogen
release in January 2019, by the continuous operation of the cycle for
production cycle that is considered for coupling with nuclear power
150 h with hydrogen production rate of 30 L/h and successful so-
plants and expected to operate at relatively high efficiency. The
lution circulation in the HI section for three times. Continuous ef-
main merit of this cycle is that all chemicals are in either gas or
forts are ongoing to achieve sustainable long-term operation using
liquid form which reduces the power required for transporting
industrial materials, and to achieve the coupling with the HTGR
these chemicals (in contrast, most of other cycles require solid
design of JAEA, HTTR (Kubo et al., 2019).
transport). There are several types for the cycle, Fig. 3 presents a
In China, the development of the SeI has progressed nicely since
simplified process diagram for the most common SeI cycle
started with fundamental studies on the cycle in 2005. The Institute
configuration, showing its three reactions: the endothermic
of Nuclear and New Energy Technology (INET) of China has suc-
decomposition of H2SO4 (oxygen production process), the
ceeded in 2009 in conducting a closed-cycle operation of contin-
exothermic Bunsen process (hydrolysis step), and the endothermic
uous 8 h with the production of 10 L of hydrogen per hour as a
decomposition of HI (hydrogen production process); operating at
proof-of-concept and proof of the process feasibility. Later in 2013,
temperature values 800e850  C, 120  C, and 300e450  C, respec-
a laboratory scale facility, IS-100, was constructed with the aim of
tively. However, the heat source needed for SeI cycle is to be at least
producing 100 L/h, and in 2014 a continuous stable operation of the
at 950  C to count for the heat losses in the heat exchanger. The
system for over 80 h was achieved with 60 h of hydrogen pro-
reactions of the cycle consume ca. 270e375 kJ of thermal energy
duction at rate of 60 L/h (Zhang et al., 2018). This was a proof of
per mole of hydrogen production. In addition, around 50e75 kJ of
cycle controllability and stability, which is followed by continuous
electric power per mole of produced hydrogen is required to
work towards the up-scaling of the process with its coupling to the
operate the auxiliary systems between the different processes of
HTGR design of HTR-PM600 and HTR-10, and the associated safety
the cycle.
concerns, engineering materials selection, and improving operating
Several countries have their own national and international
efficiency of the cycle (Zhang et al., 2019).
collaborative projects for developing a pilot scale prototype of the
In addition to several achievements in the Republic of Korea on
cycle which was first invented in 1970s. One of the initiatives was
the process simulation and the investigation of its static and dy-
the International Nuclear Energy Research Initiative which was
namic performance, the Korean Atomic Energy Research Institute
conducted with collaboration between Sandia National
(KAERI) and the Korea Institute of Energy Research (KIER) have
constructed a 50 L/h scale test system to be operated under pres-
surized environment and the start-up dynamic behavior of the
HIeI2eH2O distillation column was studied and reported in the
literature (Shin et al., 2015).

2.2. HyS cycle

The HyS cycle, proposed in mid-1970s, was first developed by


Westinghouse electric corporation (it is also known as the West-
inghouse Cycle) (Brecher and Wu, 1975). An electrochemical step
along with the thermochemical sulfuric acid decomposition step,
hence called hybrid cycle, brings simplification to the cycle
compared to the traditional SeI cycle. This cycle requires around
180 kJ of thermal energy per mole of hydrogen along with
55e80 kJ/mol of electric power.
Further investigation followed in 1970s and 1980s in Westing-
house (Carty et al., 1977; Lu and Ammon, 1982; Lu, 1983). During
the same period, the cycle was studies in Europe, including
research conducted at the European Joint Research Centre and the
German Nuclear Research Centre in Jülich (Struck et al., 1980;
Fig. 3. SeI thermochemical cycle. Junginger and Struck, 1982). The HyS comprises two chemical
4 R.S. El-Emam et al. / Journal of Cleaner Production 262 (2020) 121424

reactions: sulfur dioxide depolarized electrolysis of water and


sulfuric acid decomposition. A simplified diagram of the cycle two
steps is shown in Fig. 4.
The thermochemical SO2 depolarized electrolysis of water yields
sufuric acid and hydrogen.

SO2ðaqÞ þ 2H2 OðgÞ /H2 SO4ðaqÞ þ H2ðgÞ at 80  120 C

This process requires one-third of the electric power required


for direct water electrolysis (Dokyia et al., 1977) and this is what
makes HyS process interesting. The SO2 electrolyzer uses the con-
ventional polymer electrolyte membrane (PEM) electrolysis.
However, the depolarization of the anode by SO2 reduces the
standard 1.23 V cell potential. The theoretical potential of this
process is 0.158 V, yet it is targeted to be brought down to less than
half of this value at 500 mA/cm2. This process actually occurs in two
main steps. First, the electrochemical oxidization of SO2 at the
anode forming sulfuric acid, protons and electrons.

SO2ðaqÞ þ 2H2 OðgÞ /H2 SO4ðaqÞ þ 2Hþ þ 2e

The protons are then conducted through the electrolyte sepa-


rator to the cathode side to form hydrogen with the electrons.

2Hþ þ 2e /H2ðgÞ

Different designs of the SO2-depolarized electrolysis were


developed after the initial design of Westinghouse. Fig. 5 shows
two schematics of different SO2-depolatized PEM electrolysis
designs.
The endothermic thermochemical reaction of sulfuric acid
decomposition requires heat at temperature of at least 750  C.

H2 SO4ðaqÞ / SO2ðgÞ þ H2 OðgÞ þ 0:5O2ðgÞ at 800 C

This thermochemical reaction step of the HyS cycle is similar to


the one occurring at the SeI process and is common in all other
sulfur cycles. Sulfuric acid is dehydrated through heating and
eventually sulfur dioxide is formed and then it thermally de-
composes. The decomposition of sulfuric acid into SO3 and vapor
Fig. 5. Schematic of SO2 depolarized PEM electrolysis cell with water fed at the
occurs first at around 800  C or more as bellow: cathode and the anode is fed with: (a) SO2 dissolved in sulfuric acid and (b) gaseous
SO2.
H2 SO4ðaqÞ / SO3 þ H2 OðgÞ

This is followed by the catalytic decomposition of the resulted


SO3 producing sulfur dioxide and oxygen. This part of the decom- SO3 / SO2ðgÞ þ 0:5O2ðgÞ
position is processed at lower temperature, yet higher than 450  C.
Different alternatives of the cycle were investigated in several
countries, including: USA, France, Japan, and Canada, to reach a
configuration with lower maximum temperature for the endo-
thermic reaction, considering heat source of 500e700  C. Other
projects were also conducted in the Republic of Korea and South
Africa. Along with the SO2 transportation through the electrolyzer
membrane and the sulfur buildup at certain parts of the cycle,
finding the proper catalyst for the electrolysis to proceed is one of
the main challenges for this cycle.
On a related topic, the use of sulfur based thermochemical en-
ergy storage is considered by the ongoing PEGASUS project funded
by Horizon 2020 European Union programme. Concentrated solar
thermal energy is utilized for the decomposition of sulfuric acid
where the produced sulfur dioxide is decomposed in a dispropor-
tionation reaction to elemental sulfur and sulfuric acid. Sulfur is
later combusted in air producing high quality heat of over 1000  C.
Even though this is not a hydrogen production route, yet the
development of this part of the cycle at industrial large-scale level
would help enabling sulfur-based thermochemical cycle. Anyway,
Fig. 4. HyS hybrid thermochemical cycle. the decomposition of sulfuric acid has been already investigated in
R.S. El-Emam et al. / Journal of Cleaner Production 262 (2020) 121424 5

the literature. In general, the HyS cycle faces some challenges


including the optimization of the electrochemical step of the cycle
(i.e. sulfur dioxide depolarization of electrolysis) reducing the
process overpotential. Another technical challenge in this step is
sulfur dioxide transport through the membrane. The catalysts ac-
tivity and degradation in the electrolyzer is another challenge that
has been widely investigated in the literature and still under
development (Diaz-Abad et al., 2019), along with the investigation
on high-corrosion resistive materials for the cycle processes in
general. The cycle should be avoided to operate below 120  C to
avoid sulfur buildup in the cycle components as this is the nominal
melting point of sulfur.

2.3. CueCl cycle


Fig. 7. Hybrid thermochemical CueCl cycle with hydrogen evolving electrochemical
Various configurations of the hybrid copper-chlorine cycle were step.
proposed starting with the work of Dokyia and Kotera (1976) and
the US Institute of Gas Technology as reported in Carty et al. (1981).
In one main variant, the electrochemical process of this hybrid cycle electrochemical unit. Hydrogen evolves separately, at the cathodic
generates copper from CuCl oxidation. This variant is known as the circuit of the electrochemical cell which include a membrane for
five-step CueCl hybrid thermochemical cycle and its main pro- separation of anolyte from catholyte (cathodic electrolyte). Cath-
cesses are shown in Fig. 6. Hydrogen is generated in a thermo- olyte typically consists of a strong solution of HCl(aq).
chemical reaction of chlorination, and oxygen is produced by a From Zamfirescu et al. (2017) a flow sheet of CueCl cycle con-
thermolysis process. sisting of three reaction steps and one separation step is shown in
Starting with the work published in Naterer et al. (2010) a more Fig. 8. The separation step is complex. It is driven thermochemi-
appealing four step version of the copper-chlorine cycle, which cally, and also involves pumps, stirrers and conveyors. The sepa-
includes a hydrogen-generating electrochemical step is in devel- ration process involves drying with recycling of strong HCl(aq)
opment. The four-step cycle is based on two thermochemical pro- though subsequent condensation. Drying produces a mixed salt
cesses, one electrochemical step and a separation process. In consisting of CuCl and CuCl2. Subsequent separation of salts is
Naterer et al. (2010) the separation process in based on CuCl2 conducted to a selective dissolution of CuCl2 in water, as CuCl re-
crystallization in an aqueous solution, but in other developments a mains insoluble and can be extracted from sediment or by filtration.
drying process is used to extract salts from the electrolyte oxidized On additional separation aspect relates to the making-up of HCl to
in the electrochemical cell. In Naterer and Wang (2016) a drier- the required concentration for the reduced anolyte. In electro-
based four-step copper-chlorine thermochemical water splitting chemical unit HCl is consumed. To make up the HCl, the output of
cycle is disclosed. the hydrolysis unit is used, which generates HCl. Nevertheless, no
Zamfirescu et al. (2017) recognizes the fundamental role of the pure HCl is obtained anywhere within the cycle. HCl is present
anodic electrolyte e also known as anolyte e in chemical transport, either as mixed vapor with steam or as aqueous solution. Water
separations, and chemical conversions within the cycle. In a must be then removed from the oxidized anolyte, while in the same
reduced form, the anolyte consists of CuCl dissolved in strong time strong HCl(aq) is added to re-make the reduced anolyte. A
HCl(aq) and is feed at the electrochemical process. In the oxidized pressure swing distillation unit to concentrate HCl(aq) and recycle
form, anolyte consists of CuCl and CuCl2 dissolved in strong HCl(aq) water is therefore one recommended choice. This unit can be
as it is generated in the electrochemical unit. Fig. 7 shows how operated on thermal energy at max. 200  C. Making-up the anolyte
anolyte bridges between electrochemical and thermochemical also requires addition of CuCl from thermolysis, which yield as
processes of the cycle. As shown in El-Emam et al. (2019), the cycle molten salt, but is further quenched in water and separated as
requires ca. 200 kJ per mole of H2 to run its thermochemical pro- dense sediment. The reduced anolyte is produced in the dissolution
cesses, which split water molecule and generate oxygen gas, while cell of recycled and newly generated CuCl; where the salt dissolved
in the same time reduce the anolyte which has been oxidized in the in strong HCl(aq). A detailed flowsheet of the four-step cycle pre-
viously discussed is due to Farsi et al. (2019) .

2.4. MgeCl cycle

MgeCl cycle is a configuration of the Reverse Deacon cycle in


which HCl electrolysis for H2 and Cl2 production occurs. Here HCl
production is accomplished with subsequent reactions of MGCl2 hy-
drolysis and MgO chlorination (Simpson et al., 2006). The three-step
cycle has potential problems of having high maximum temperature
and intense heat requirement for hydrolysis while the pore size of the
produced MgO particles do not allow chlorination at higher reaction
rates (Ng et al., 2005; Kashani-Nejad et al., 2005). Therefore, an
additional step is developed based on the two known three-step
configurations of this cycle to decrease cycle thermal and electrical
energy requirements (Ozcan and Dincer, 2016a, Ozcan and Dincer,
2018a,b). The four-step cycle shown in Fig. 9 considers an interme-
Fig. 6. Five-Step CueCl hybrid thermochemical cycle. diate exothermic reaction for MGCl2 hydrolysis and then decomposes
6 R.S. El-Emam et al. / Journal of Cleaner Production 262 (2020) 121424

Fig. 8. Flowsheet of a four-step CueCl water splitting cycle, based on Zamfirescu et al. (2017).

2.5. CaeBr cycle

A four-step cycle has first been proposed and analyzed as a pure


thermochemical form called the UT-3 cycle, where hydrogen pro-
duction step requires use of iron-oxides (Nakayama et al., 1984).
The UT-3 cycle uses iron bromide and iron oxide intermediates for
recycling the reagents and for hydrogen evolving in two reaction
steps (Yoshida et al., 1990; Sakurai et al., 1996b; Simpson et al.,
2007; Teo et al., 2005). The UT-3 reactions are as follows:

CaBr2ðsÞ þ H2 OðgÞ /CaOðsÞ þ 2HBrðgÞ at 700  750 C

3FeBr2ðsÞ þ 4H2 OðgÞ /Fe3 O4ðsÞ þ 6HBrðgÞ þ H2ðgÞ at 550



 600 C

Fe3 O4ðsÞ þ 8HBrðgÞ /3FeBr2ðsÞ þ Br2ðgÞ þ 4H2 OðgÞ at 200



 300 C
Fig. 9. MgeCl hybrid thermochemical cycle.

CaOðsÞ þ Br2ðgÞ /0:5O2ðgÞ þ CaBr2ðsÞ at 500  600 C

the product (MgOHCl) for dry HCl production at 450  C. This config- In this cycle, solid reactants are used cyclically in heterogeneous
uration solves two main problems of the MgeCl cycle. First, MgO reactors. The reaction rate of CaBr2 hydrolysis reaction is the
produced from MgOHCl decomposition has higher reactivity with the slowest, which limits the rates of other reactions. A hybrid version
chlorine gas and second, dry HCl electrolysis requires lower electrical of the UT-3 cycle was proposed, using a single electrochemical re-
energy consumption. Comparative thermal analysis and integration action of dissociation of hydrobromic acid at 50  C, which evolves
of the four-step cycle can be found elsewhere (Ozcan and Dincer, hydrogen and regenerates bromine.
2016b; Ozcan and Dincer, 2016c). Mature chemical reactions and The hybrid CaeBr cycle is a modified version of the UT-3 cycle
electrolysis step, and favorable reaction rates makes the cycle a which includes four-reaction steps with heat requirements at
feasible one for further consideration (Ozcan, 2015). The cycle theo- around 750  C and up to 800  C (Doctor et al., 2004; Lottes et al.,
retically consumes around 151.5 kJ of thermal energy and 190 kJ of 2009; Aihara et al., 1990; Sakurai et al., 1996a). This hybrid imple-
electric power per mole of produced hydrogen with theoretical cell mentation of the CaeBr in an electro-thermochemical cycle con-
potential of 0.99 V. siders two thermochemical reaction in addition to HBr electrolysis
R.S. El-Emam et al. / Journal of Cleaner Production 262 (2020) 121424 7

occurring at temperature not more than 100  C. These are two gas- research is required for the reaction kinetics which may lead to
solid reactions and one gas phase reaction: increased maximum cycle temperatures. It is depicted that the
maximum cycle temperature is achievable at around 500  C at
CaBr2ðsÞ þ H2 OðgÞ /CaOðsÞ þ 2HBrðgÞ at 750 C vacuum pressure conditions which in general accepted as a chal-
lenging process as depicted in previous studies for other promising
CaOðsÞ þ Br2ðgÞ /0:5O2ðgÞ þ CaBr2ðsÞ at 580 C cycles (Marques et al., 2018). An exergy analysis for the cycle was
found as 83% while practical aspects are not included such as excess
water ratio (Marques et al., 2020).
2HBrðgÞ / Br2ðgÞ þ H2ðgÞ at 100 C

The decomposition of HBr through electrolysis in this cycle 3. Economics, environmental and safety aspects
replaced the other versions using thermal decomposition and
plasma reaction. However, the main challenge in this configuration As discussed, due to its abundance, steadiness, and trans-
was due to the low solubility of CaO in CaBr2 and deposition of CaO portability, nuclear energy is a feasible option for clean and cost-
at the reactor inlet (Simpson et al., 2007). This three-step cycle effective hydrogen production. However, many aspects need to be
phases out the use of iron-oxide and directly electrolyzes CaeBr taken into account starting from economics to stakeholders. Clean
agent from the hydrolysis step which makes the cycle simpler. hydrogen production using conventional electrolysis might be the
However, electricity is required to drive the process. The four-step most cost-effective route when the driving electricity is generated
cycle efficiency is as high as 50% while the hybrid three-step cycle from nuclear or geothermal energy systems (El-Emam and Ozcan,
may possess around 47% energy efficiency with the electrolysis 2019). Even though thermochemical cycles are designed for the
process that requires less than half of the electrical energy required production of low-cost hydrogen efficiently, the need for high
for conventional water electrolysis (Sakurai et al., 1996b). The temperature and material compatibility issues bring high cost
maximum cycle temperature makes this hybrid cycle compatible associated with technology and safety needs. Fig. 11 represents cost
for coupling with advanced nuclear reactors. Even though many ranges of nuclear hydrogen generation using different hydrogen
previous studies have been conducted for the pure CaeBr cycle, not production technologies. It should be noted that the variations in
many detailed studies are present for the hybrid version. Fig. 10 the costs are dependent on capacity, size, and other specific plant
shows a schematic of the hybrid version of the cycle. configuration. In the figure, one can see that HyS brings a wide
range of hydrogen production cost variation compared to other
2.6. Other promising cycles technologies since many studies are present for stoichiometric and
lab scale studies. Reported cost for the CueCl and MgeCl cycles are
Even though experimental results show reasonable reaction among the lowest when considering stoichiometric conditions
kinetics for the two-step pure Zn/Zno and Cr/CrO thermochemical while high limits of their cost ranges are for more practical sce-
cycles, very high temperature requirements make these configu- narios. Under realistic conditions, the ratio of steam for hydrolysis
rations only compatible with concentrated solar energy systems. is very large and this requires larger components and reactors for
The U-Eu-Br heavy-element halide cycle has been proposed as a the cyclic process to complete, hence, higher plant investment
pure thermochemical cycle that can even match with the existing costs. HTSE is of relatively lower cost compared to thermochemical
nuclear reactors that has a maximum temperature of 310  C only, cycles as it does not require large reactors nor controlled cyclic
makes this cycle to be the lowest temperature requiring pure processes. The highest cost contributor for the HTSE cycle is the fuel
thermochemical cycle (Petri et al., 2007). The proof-of-concept cost (i.e. cost of electric power) even though the electricity con-
reactions show proceed in reactions in very small experimenta- sumption is over 20% less than that of low temperature electrolysis.
tion. Further studies are required for reactor integration, thermal There are several tools used for hydrogen cost estimation such as
analysis and reaction kinetics, as well as for safety. Hydrogen from Generation IV Excel Calculations on Nuclear Systems (G4ECONS)
low temperature sodium redox reactions is first proposed to be and Hydrogen Economy Evaluation Program (HEEP), mainly using
accomplished at 400  C (Miyaoka et al., 2012). Individual reactions nuclear energy (Sadhankar et al., 2018; El-Emam et al., 2015; Ozcan
are designed as cyclic processes to represent their feasibility. Even et al., 2014; El-Emam and Khamis, 2017, 2019). US Department of
though the cycle represents low maximum temperatures, further Energy (DOE) targets the centralized hydrogen cost from electrol-
ysis to be $2/kg by 2020. This target probably requires govern-
mental promotions to encourage investment.
With regards to environmental aspects, two of the indicators for
environmental impact assessment are considered here: Global

Fig. 11. Cost of hydrogen produced using nuclear energy driven technologies (Modified
Fig. 10. CaeBr hybrid thermochemical cycle. from El-Emam and Ozcan, 2019).
8 R.S. El-Emam et al. / Journal of Cleaner Production 262 (2020) 121424

Warming Potential (GWPeCO2eq) and Acidification Potential larger markets of energy storage, transportation and power gen-
(APeSO2eq). Fuel mining, preparation, transportation and use eration. Nuclear energy is one of the main alternatives to cope with
include a major amount of total emissions from a nuclear plant. the thermal and electrical energy demand for the promising
Even though there are also emissions due to construction and hydrogen production technologies considered for large-scale pro-
operation of nuclear plants, their environmental impacts are duction of low/zero-carbon hydrogen. Some of the pure and hybrid
significantly lower than conventional fossil-driven power plants thermochemical technologies are among the promising technolo-
that makes nuclear source a clean and sustainable one. Hydrogen gies considered for nuclear hydrogen production.
generation plant is also of importance when it comes to environ- The SeI cycle shows good potential for large-scale hydrogen
mental impact. Requirement for noble materials and use of mate- production at competitive cost and high efficiency. However, it
rials of finite nature in electrolysis processes have impact on the faces some chemical kinetics and solution thermodynamics un-
environment. Hybrid and pure thermochemical cycles have lower certainties that were discussed in the literature and are still being
environmental impact than that of nuclear driven HTSE and PEM investigated. Among the challenges facing this cycle are the
plants (El-Emam and Ozcan, 2019). Nuclear driven hydrogen pro- worldwide availability of iodine for supporting large-scale sus-
duction is one most environmentally benign options for clean tainable hydrogen production and the near-term availability of
hydrogen generation that shows promising values with hydro- and economic and clean sustainable heat source. With regards to the
wind-driven hydrogen technologies, and significantly lower than HyS cycle, the main elements requiring further development to
those of biomass gasification and Photovoltaic based reach commercial scale include: improving electrocatalysts to
configurations. reduce the cell potential of the SO2 depolarized water electrolysis,
When it comes to safety aspects, hydrogen carry a large range of resolving the problem of SO2 crossover during the process, and
flammability which makes it risky compared to other explosives long-term stable operation at low operating potential degradation.
even though it has lower flammability limit than many other The target operating condition of the electrolysis process of the HyS
combustibles (Fischer, 1986). However, this does not provide any cycle is around 0.6 V per cell at around 120  C or less. This gives
advantage over other combustibles since ignition sources with the potential to bring electricity consumption to one quarter that of
lower limit is always present. Hydrogen also has higher burning direct water electrolysis. The CueCl cycle has seen recent progress
velocity and similar flame temperature with compared gases. in the development on integrated laboratory scale demonstration.
Deflagration related overpressures of hydrogen-air mixture under Further steps towards a pilot plant would be necessary before this
stoichiometric conditions are at a maximum of 8:1 while it doubles cycle can take a solid step towards commercialization. Other cycles,
when the detonation occurs in closed containments such as pipe- including MgeCl and CaeBr cycles have been seeing slow progress
lines, which is the leading cause for deflagration-to-detonation and some are still facing large amount of uncertainties and chal-
transition (DDT) (Crowl, 1992). The theoretical energy content of lenges with regards to the chemical kinetics, thermodynamics of
hydrogen is 24 times higher than the same mass of TNT, which involved solutions, and chemistry of occurring side reactions.
provides a theoretical upper limit for explosion energy. However, In general, the availability of low-carbon cost-effective heat
low energy density of hydrogen decreases this value to 2, under source at the necessary quality (i.e. temperature) required for the
atmospheric conditions. Some experimental results suggested that corresponding hydrogen thermochemical cycle is possible consid-
only 10% of the theoretical energy is released because only small ering the advanced nuclear reactor designs. However, the current
amount of released hydrogen is at flammability range (Crowl, 1992, fleet of nuclear power plants as well as those under construction
Jeremi c and Baji
c, 2006). Correlation of survivability after being still do not meet the requirements to be coupled with medium/high
exposed to overpressures due to explosion has been conducted by temperature thermochemical cycles of hydrogen production. The
Gibson (1994) that provides an adequate insight on possible effect consideration of heat-quality upgrade may facilitate such
of hydrogen explosion on people. Various sizes of TNT explosives integration.
are considered for the correlations. Specified effect of a blast on
people and structures are based on phase duration of an explosive
and overpressure related shock waves (Tolias et al., 2018). References
Studies on hydrogen safety has increased after feasible tech-
Aihara, M., Umida, H., Tsutsumi, A., Yoshida, K., 1990. Kinetic study of UT-3 ther-
nologies are demonstrated. A report based on experiments mochemical hydrogen production process. Int. J. Hydrogen Energy 15 (1), 7e11.
demonstrated that hydrogen can be safer than other gaseous Brecher LE. Wu CK. Electrolytic decomposition of water. US Patent 3,888,750:
combustibles based on its low energy density and fast dissipation Westinghouse Electric Corp., June 10, 1975.
Carty, R., Cox, K., Funk, J., Soliman, M., Conger, W., Brecher, L., Spewock, S., 1977.
while less safe due to its tendency to detonate and wide flamma- Process sensitivity studies of the Westinghouse sulfur cycle for hydrogen gen-
bility range (Cracknell et al., 2002). Review of many hydrogen ex- eration. Int. J. Hydrogen Energy 2 (1), 17e22.
plosion accidents in the world provide suggestions for safer Carty, R.H., Mazumder, M.M., Schreiber, J.D., Pangborn, J.B., 1981. Thermochemical
Production of Hydrogen. Final report 30517, vols. 1e4. Institute of Gas Tech-
hydrogen generation by depicting on advanced ventilation in nology, Chicago, IL.
congested areas (Ng and Lee, 2008). Rodionov et al. (2011) evalu- Cracknell, R.F., Alcock, J.L., Rowson, J.J., Shirvill, L.C., Üngüt, A., 2002 Jan 1. Safety
ated risks of hydrogen use in mobile applications and they resulted considerations in retailing hydrogen. SAE Trans. 922e926.
Crowl, D.A., 1992 Jan 1. Calculating the energy of explosion using thermodynamic
that added risk of hydrogen use is only 0.2% which is a residual availability. J. Loss Prev. Process. Ind. 5 (2), 109e118.
contribution. Verfondern et al. (2017) determined safe distances of Diaz-Abad, S., Millan, M., Rodrigo, M.A., Lobato, J., 2019. Review of anodic catalysts
hydrogen plants from the nuclear reactor when nuclear based for SO2 depolarized electrolysis for “green hydrogen” production. Catalysts 9,
63. https://doi.org/10.3390/catal9010063.
hydrogen generation is considered. Risks of toxic leakages from the Doctor, R.D., Matonis, D.T., Wade, D.C., Moisseytsev, A.V., Sienicki, J.J., Faibish, R.,
Sulfur-Iodine hydrogen plant are also reported. More information is 2004. STAR-H2 with a calcium bromine cycle: delivering hydrogen, electricity
available at El-Emam and Dincer (2018). and water from a modular reactor. In: Proceedings of the Spring National
Meeting of the American Institute of Chemical Engineers. April 25e29.
Dokyia, M., Fukuda, K., Kameyama, T., Kotera, Y., Asakura, S., 1977. The study of
4. Conclusion thermochemical hydrogen preparation. (II) Electrochemical hybrid cycle using
sulphur-iodine system. Denki Kagaku 45, 139e143.
Today, hydrogen is widely used in several industries including Dokyia, M., Kotera, Y., 1976. Hybrid cycle with electrolysis using Cu-Cl system. Int. J.
Hydrogen Energy 1, 117e121.
the production of fertilizers and in oil refineries. This is expected to El-Emam, R.S., Dincer, I., Zamfirescu, C., 2019. Enhanced CANDU reactor with heat
expand much more in the near future for serving new sectors and upgrade for combined power and hydrogen production. Int. J. Hydrogen Energy
R.S. El-Emam et al. / Journal of Cleaner Production 262 (2020) 121424 9

44 (42), 23580e23588. Ng, K.W., Kashani-Nejad, S., Harris, R., 2005 Jun 1. Kinetics of MgO chlorination with
El-Emam, R.S., Dincer, I., 2018. Nuclear-assisted hydrogen production. In: Meyers, R. HCl gas. Metall. Mater. Trans. B 36 (3), 405e409.
(Ed.), Encyclopedia of Sustainability Science and Technology. Springer, New Noguchi, H., Takegami, H., Kasahara, S., Tanaka, N., Kamiji, Y., Iwatsuki, J., Aita, H.,
York, NY. Kubo, S., 2017. R&D status in thermochemical water-splitting hydrogen pro-
El-Emam, R.S., Khamis, I., 2019. Advances in nuclear hydrogen production: results duction iodine-sulfur process at JAEA. Energy Procedia 131, 113e118.
from an IAEA international collaborative research project. Int. J. Hydrogen En- Ozcan, H., 2015. Experimental and Theoretical Investigations of Magnesium-
ergy 44 (35), 19080e19088. Chlorine Cycle and its Integrated Systems. PhD Thesis. University of Ontario
El-Emam, R.S., Khamis, I., 2017. International collaboration in the IAEA nuclear Institute of Technology.
hydrogen production program for benchmarking of HEEP. Int. J. Hydrogen En- Ozcan, H., Dincer, I., 2016a. Comparative performance assessment of three config-
ergy 42 (6), 3566e3571. urations of magnesiumechlorine cycle. Int. J. Hydrogen Energy 41 (2), 845e856.
El-Emam, R.S., Ozcan, H., Dincer, I., 2015. Comparative cost evaluation of nuclear Ozcan, H., Dincer, I., 2018a. Energetic and exergetic performance comparisons of
hydrogen production methods with the Hydrogen Economy Evaluation Pro- various flow sheet options of magnesium-chlorine cycle. In: Exergy for A Better
gram (HEEP). Int. J. Hydrogen Energy 40 (34), 11168e11177. Environment and Improved Sustainability, vol. 1. Springer, Cham,

El-Emam, R.S., Ozcan, H., 2019. Comprehensive review on the techno-economics of pp. 1191e1206.
sustainable large-scale clean hydrogen production. J. Clean. Prod. 220, Ozcan, H., Dincer, I., 2018b. Experimental investigation of an improved version of
593e609. the four-step magnesium-chlorine cycle. Int. J. Hydrogen Energy 43 (11),
Farsi, A., Zamfirescu, C., Dincer, I., Naterer, G.F., 2019. Thermodynamic assessment of 5808e5819.
a lab-scale experimental copper-chlorine cycle for sustainable hydrogen pro- Ozcan, H., Dincer, I., 2016b. Modeling of a new four-step magnesiumechlorine cycle
duction. Int. J. Hydrogen Energy 44, 17595e17610. with dry HCl capture for more efficient hydrogen production. Int. J. Hydrogen
Fischer, M., 1986 Jan 1. Safety aspects of hydrogen combustion in hydrogen energy Energy 41 (19), 7792e7801.
systems. Int. J. Hydrogen Energy 11 (9), 593e601. Ozcan, H., Dincer, I., 2016c. Thermodynamic modeling of a nuclear energy based
Gibson, P., 1994. Blast Overpressure and Survivability Calculations for Various Sizes integrated system for hydrogen production and liquefaction. Comput. Chem.
of Explosive Charges. Army Natick Research Development and Engineering Eng. 90, 234e246.
Center MA. Ozcan, H., El-Emam, R.S., Dincer, I., 2014. Comparative assessment of nuclear based
IAEA, 2019. PRIS-STATISTICS: Power Reactor Information System Statistical Reports, hybrid sulfur cycle and high temperature steam electrolysis systems using
Computer Manual Series No. 22. The International Atomic Energy Agency, HEEP. In: Progress in Sustainable Energy Technologies: Generating Renewable
Vienna. Energy. Springer, Cham, pp. 165e180.
IEA, 2019a. The Future of Hydrogen: Seizing Today’s Opportunities. The Interna- Petri, M.C., Klickman, A.E., Hori, M., 2007. Hydrogen production options for water-
tional Energy Agency, Technology Report, Paris. June. cooled nuclear power plants. In: International Conference on Non-electric
IEA, 2019b. Nuclear Power in a Clean Energy System. The International Energy Applications of Nuclear Power, Oarai, Japan.
Agency, Fuel Report, Paris. Rodionov, A., Wilkening, H., Moretto, P., 2011. Risk assessment of hydrogen explo-
Jeremi c, R., Baji
c, Z., 2006. An approach to determining the TNT equivalent of high sion for private car with hydrogen-driven engine. Int. J. Hydrogen Energy 36 (3),
explosives. Sci. Technol. Rev. 56 (1), 58e62. 2398e2406.
Junginger, R., Struck, B.D., 1982. Separators for electrolytic cells of the sulfuric acid Sadhankar, R., Sopczak, L., Ryland, D., El-Emam, R.S., Khamis, I., 2018. In: Bench-
hybrid cycle. Int. J. Hydrogen Energy 7 (4), 331e340. marking of economic models for nuclear hydrogen production, Advancing and
Kashani-Nejad, S., Ng, K.W., Harris, R., 2005 Feb 1. MgOHCl thermal decomposition sustaining nuclear energy, vol. 2: Pacific basin nuclear conference (PBNC 2018),
kinetics. Metall. Mater. Trans. B 36 (1), 153e157. 30 September - 4 October 2018, San Francisco, California, USA, pp. 889e897.
Kubo, S., Nakajima, H., Kasahara, S., Higashi, S., Masaki, T., Abe, H., Onuki, K., 2004. Sakurai, M., Bilgen, E., Tsutsumi, A., Yoshida, K., 1996b. Adiabatic UT-3 thermo-
A demonstration study on a closed-cycle hydrogen production by the ther- chemical process for hydrogen production. Int. J. Hydrogen Energy 21 (10),
mochemical water-splitting iodine-sulfur process. Nucl. Eng. Des. 233, 865e870.
347e354. Sakurai, M., Bilgen, E., Tsutsumi, A., Yoshida, K., 1996a. Solar UT-3 thermochemical
Kubo, S., Takegami, H., Tanaka, N., Noguchi, H., Kamiji, Y., Jin, Iwatsuki, cycle for hydrogen production. Sol. Energy 57 (1), 51e58.
Myagmarjav, Odtsetseg, Inagaki, Yoshiyuki, 2019. Massive and efficient H2 Shin, Y., Lee, T., Lee, K., 2015. Start-up dynamic behaviors of a H-I2-H2O distillation
production technology on thermochemical water-splitting iodine-sulfur pro- column for the sulfur-iodine hydrogen production cycle. Int. J. Hydrogen Energy
cess. In: Presented at the World Hydrogen Technology Conference, Tokyo, 40 (39), 13264e13271.
Japan. Simpson, M.F., Herrmann, S.D., Boyle, B.D., 2006. A hybrid thermochemical elec-
Lottes, S.A., Lyczkowski, R.W., Panchal, C.B., Doctor, R.D., 2009. Modeling and trolytic process for hydrogen production based on the reverse Deacon reaction.
analysis of calcium bromide hydrolysis. Int. J. Hydrogen Energy 34 (9), Int. J. Hydrogen Energy 31 (9), 1241e1246.
4155e4167. Simpson, M.F., Utgikar, V., Sachdev, P., McGrady, C., 2007. A novel method for
Lu, P.W.T., Ammon, R.L., 1982. Sulfur dioxide depolarized electrolysis for hydrogen producing hydrogen based on the CaeBr cycle. Int. J. Hydrogen Energy 32 (4),
production: development status. Int. J. Hydrogen Energy 7 (7), 563e575. 505e509.
Lu, P.W.T., 1983. Technological aspects of sulfur dioxide depolarized electrolysis for Steinfeld, A., 2002 Jun 1. Solar hydrogen production via a two-step water-splitting
hydrogen production. Int. J. Hydrogen Energy 8 (10), 773e781. thermochemical cycle based on Zn/ZnO redox reactions. Int. J. Hydrogen Energy
Marques, J.G., Costa, A.L., Pereira, C., 2018. NaOH thermochemical water splitting 27 (6), 611e619.
cycle: a new approach in hydrogen production based on sodium cooled fast Struck, B.D., Junginger, R., Boltersdorf, D., Gehrmann, J., 1980. The anodic oxidation
reactor. Int. J. Hydrogen Energy 43 (16), 7738e7753. of sulfur dioxide in the sulfuric acid hybrid cycle. Int. J. Hydrogen Energy 5 (5),
Marques, J.G., Costa, A.L., Pereira, C., 2020. Exergy analysis for the Na-OH (sodium- 487e497.
oxygen-hydrogen) thermochemical water splitting cycle. Int. J. Hydrogen Teo, E.D., Brandon, N.P., Vos, E., Kramer, G.J., 2005. A critical pathway energy effi-
Energy. ciency analysis of the thermochemical UT-3 cycle. Int. J. Hydrogen Energy 30,
Miyaoka, H., Ichikawa, T., Nakamura, N., Kojima, Y., 2012. Low-temperature water- 559e564.
splitting by sodium redox reaction. Int. J. Hydrogen Energy 37 (23), Tolias, I.C., Giannissi, S.G., Venetsanos, A.G., Keenan, J., Shentsov, V., Makarov, D.,
17709e17714. Coldrick, S., Kotchourko, A., Ren, K., Jedicke, O., Melideo, D., 2018. Best practice
Nakajima, H., Ikenoya, K., Onuki, K., Shimizu, S., 1998. Closed-cycle continuous guidelines in numerical simulations and CFD benchmarking for hydrogen safety
hydrogen production test by thermochemical IS process. Kagaku Kogaku Ron- applications. Int. J. Hydrogen Energy.
bunshu 24, 352e355. Verfondern, K., Yan, X., Nishihara, T., Allelein, H.J., 2017. Safety concept of nuclear
Nakayama, T., Yoshioka, H., Furutani, H., Kameyama, H., Yoshida, K., 1984 Jan 1. cogeneration of hydrogen and electricity. Int. J. Hydrogen Energy 42 (11),
MASCOTda bench-scale plant for producing hydrogen by the UT-3 thermo- 7551e7559.
chemical decomposition cycle. Int. J. Hydrogen Energy 9 (3), 187e190. Yoshida, K., Kameyama, H., Aochi, T., Nobue, M., Aihara, M., Amir, R., et al., 1990.
Naterer G., Wang Z. 2016. Systems, Methods and Devices for the Capture and Hy- A simulation study of the UT-3 thermochemical hydrogen production process.
drogenation of Carbon Dioxide with the Thermochemical Cu-Cl and Mg-Cl-Na/ Int. J. Hydrogen Energy 15.
K-CO2 Cycles. US Patent 9464010 B2. Zamfirescu, C., Naterer, G.F., Rosen, M.A., 2017. Chemical exergy of electrochemical
Naterer, G.F., Suppiah, S., Stolberg, L., Lewis, M., Wang, Z., Daggupati, V., Gabriel, K., cell anolyte of cupric/cuprous chlorides. Int. J. Hydrogen Energy 42,
Dincer, I., Rosen, M.A., Spekkens, P., Lvov, L., Fowler, M., Tremaine, P., 10911e10924.
Mostaghimi, J., Easton, E.B., Trevani, L., Rizvi, G., Ikeda, B.M., Kaye, M.H., Lu, L., Zhang, P., Chen, S., Wang, L., Xu, J., 2019. R&D progress of nuclear hydrogen pro-
Pioro, I., Smith, W.R., Secnik, E., Jiang, J., Avsec, J., 2010. Canada’s program on duction in China. In: Presented at the World Hydrogen Technology Conference,
nuclear hydrogen production and the thermochemical Cu-Cl cycle. Int. J. Tokyo, Japan.
Hydrogen Energy 35, 10905e10926. Zhang, P., Wang, L., Chen, S., Xu, J., 2018. Progress of nuclear hydrogen production
Ng, H.D., Lee, J.H., 2008 Mar 1. Comments on explosion problems for hydrogen through the iodine-sulfur process in China. Renew. Sustain. Energy Rev. 81,
safety. J. Loss Prev. Process. Ind. 21 (2), 136e146. 1802e1812.

You might also like