Vous êtes sur la page 1sur 15

INTERNATIONAL JOURNAL OF ENERGY RESEARCH

Int. J. Energy Res. 2011; 35:123137


Published online 18 November 2010 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/er.1769
Potential thermochemical and hybrid cycles for
nuclear-based hydrogen production
Ibrahim Dincer
1,,
and M. Tolga Balta
2
1
Faculty of Engineering and Applied Science, University of Ontario Institute of Technology (UOIT), 2000 Simcoe Street North,
Oshawa, ON, Canada L1H 7K4
2
Department of Mechanical Engineering, Faculty of Engineering, Ege University, 35100 Bornova, Izmir, Turkey
SUMMARY
This paper discusses some potential low-temperature thermochemical and hybrid cycles for nuclear-based
hydrogen production and considers them as a sustainable option for hydrogen production using nuclear process/
waste heat and off-peak electricity. We also assess their thermodynamic performance through energy and exergy
efciencies. The results show that these cycles have good potential and become attractive due to their high overall
efciencies 50% based on a complete reaction approach. The copperchlorine cycle appears to be a highly
promising cycle for nuclear-based hydrogen production in this regard. Copyright r2010 John Wiley & Sons, Ltd.
KEY WORDS
energy; exergy; efciency; nuclear; hydrogen; thermochemical cycles
Correspondence
*Ibrahim Dincer, Faculty of Engineering and Applied Science, University of Ontario Institute of Technology (UOIT), 2000 Simcoe
Street North, Oshawa, ON, Canada L1H 7K4.

E-mail: Ibrahim.Dincer@uoit.ca
Contract/grant sponsor: Ontario Research Excellence Fund
Contract/grant sponsor: Natural Sciences and Engineering Research Council of Canada
Received 23 February 2009; Revised 15 April 2010; Accepted 8 June 2010
1. INTRODUCTION
Energy is a mainstay of an industrial society. It is,
therefore, not surprising that many important organi-
zations have attempted to analyze the future need for
energy and the availability of various energy sources.
Energy consumption growth is closely linked to
population growth, although changes in life styles
and efciency improvement have a substantial inu-
ence on the per capita annual consumption. The
structure of population and the share between urban
and rural populations also affect energy demand [1].
As a result of the worldwide increasing consumption
of energy due to an increasing population and rising
living standards in less industrialized countries, the
world faces the problem of depleting energy resources
and the impairing impact of present energy consump-
tion patterns on the global climate as well as on
humanity and the environment [2].
The risk of global climate change is of great concern
to policymakers and to the public. The relation
between the energy generation sector and environmental
pollution is being carefully considered in industrialized
countries. Before executing any power generation pro-
ject, extensive and comprehensive studies are preformed
concerning the impact of such a project on the en-
vironment. Measures for decreasing climate change and
environmental pollution are considered.
Beyond fossil fuels, the mismatch between energy
consumption and energy production becomes more
obvious. Nuclear facilities produce energy at a con-
stant rate, whereas renewable energy facilities produce
energy at a variable rate. Neither type of production
matches demand. Because of daynight and seasonal
variations of sunlight, the typical capacity factor of
solar devices is 18%. (The capacity factor is the actual
energy output in a year divided by the potential energy
output if the device were operated at full capacity for
the entire period.) The capacity factor for wind is
about 35%. For renewable energy sources, the mis-
match between generation and demand is so large that
it has been estimated that if as little as 15% of the
Copyright r 2010 John Wiley & Sons, Ltd.
electricity were produced by solar or wind, there would
be limited economic incentive to obtain more energy
from such sources, even if they are free. This is because
backup power production facilities must be built to
meet demand when these renewable energy sources are
not available [3].
Research and development of clean, economic,
stable, safe, and abundant energy should be promoted
from the viewpoint of technology as a potential mea-
sure to mitigate global warming as well as for devel-
oping large and stable energy supplies. We have various
alternative energy options to fossil fuels: solar, geo-
thermal, hydropower, nuclear energy, etc. Although
available natural energy is limited due to its stability,
quality, quantity, and density, it is certain that nuclear
energy has the potential to contribute a signicant
share of energy supply and utilization. Nuclear energy
has been almost exclusively utilized for electric power
generation, but the direct utilization of nuclear thermal
energy can be used to increase energy efciency and
thereby facilitate energy savings in the near future.
Hydrogen production is one of the key technologies for
direct utilization of nuclear thermal energy [4].
Hydrogen has ideal characteristics as an energy car-
rier. It can be stored, transported with lower loss com-
pared with electricity, and used as fuel. If necessary, the
chemical energy of hydrogen can be converted to elec-
trical energy by means of fuel cells and other devices. As
it can be produced from water and, after oxidation, it
returns to water, hydrogen is clean from the viewpoint
of environmental effects. Therefore, the realization of a
hydrogen energy system where hydrogen and elec-
tricity serve as complementary secondary energy car-
riers has been considered for a long time [5].
The need for industrial hydrogen production has
increased drastically and will continue increasing even
further in the next decade. The present economic value
of all hydrogen produced worldwide is about $300
billion/year. The growth rate is about 10% per year and
expected to be doubling to 20% per year by 2010 or
beyond [4]. A key challenge facing this rapid growth is
a sustainable route to hydrogen production. Hydrogen
production for sustainable development has been stu-
died by several investigators (e.g. [612]). Dincer [11]
has outlined the key technical and environmental issues
of current hydrogen production technologies. The
technology for producing hydrogen from a variety of
resources, including renewables, is improving. Hydro-
gen can be produced as a clean fuel from the worlds
sustainable energy sources such as nuclear, solar, wind,
hydropower, biomass, and geothermal.
Thermochemical cycles were proposed in the 1970s
as an alternative and potentially more efcient method
to produce hydrogen from water. Hydrogen produc-
tion by thermochemical water splitting is a chemical
process that accomplishes the decomposition of
water into hydrogen and oxygen using only heat or, in
the case of a hybrid thermochemical process, by a
combination of heat and electricity. Thermochemical
cycles consist of series reactions, in which water is
thermally decomposed and all other chemicals are
recycled. Only heat and water are consumed. Research
in this area was prominent from the 1970s through the
early 1980s [13]. Some proof-of-principle experiments
to demonstrate the feasibility of the hydrogen pro-
duction process through such cycles have been per-
formed [13,14], and no problems associated with
technical feasibility and potential use are however
reported. Of course, there is still much to do to scale-
up these applications to a commercial level.
Recently, there have been numerous studies [1524]
reported on several technical aspects of thermo-
chemical cycles driven by the use of process heat from
nuclear and other sources. This paper aims to discuss
various potential options for nuclear-based hydrogen
production through thermochemical and hybrid cycles
In this regard; this study covers six low-temperature
thermochemical cycles driven by nuclear-based process
and waste heat at temperatures ranging from 450 to
700 K, respectively, and presents a comparative study
of these options.
2. ENERGY OPTIONS FOR NUCLEAR-
BASED HYDROGEN PRODUCTION
In this paper, a summary of four key reactor types is
briey discussed (see Reference [25] for details).
2.1. High-temperature gas-cooled reactor
(HTGR)
Hydrogen production using thermochemical cycles has
been studied during the past four decades. The thermal
energy source for these processes has been HTGRs,
which are still under development. The maximum
operating temperature of these cycles has been limited
to about 9001C, which is the available temperature
level of gas-cooled nuclear reactors. The HTGR is one
of the most suitable nuclear reactors for producing
the secondary energy carrier hydrogen owing to its
capability of producing high-temperature heat at close
to 10001C in order to provide an efcient energy
conversion.
2.2. Advanced gas reactor (AGR)
The AGR is a commercial thermal reactor that has
been built in the U.K. for electricity production in
1550 MW
th
units, with 14 units still in operation. The
AGR core consists of uranium oxide fuel pellets in
stainless-steel cladding within graphite blocks. The
graphite acts as a moderator and carbon dioxide is the
coolant. The achievable temperature of the coolant at
the reactor exit during normal operation is around
6501C. The CO
2
circulates through the core at
4.3 MPa. For future design and implementation, there
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
124 Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
is a potential to increase the operating pressure of the
AGR in order to couple it to a direct cycle supercritical
CO
2
power conversion system. The temperature of the
reactor coolant for a future design can be driven up to
7501C with new designs and analysis. This combina-
tion can enable high-efciency, economic hydrogen
production through steam electrolysis at medium
temperatures.
2.3. Advanced high-temperature reactor
(AHTR)
The molten-saltcooled AHTR is a new reactor
concept designed to provide very high-temperature
(75010001C) heat to enable efcient low-cost thermo-
chemical production of H
2
or production of electricity
[26]. The AHTR uses the solid-coated particle fuel in
a graphite-matrix like the modular helium reactor
(MHR), but a molten-uoride-salt as coolant. It
combines the high-temperature fuel from the HTGR
with a denser coolant for the molten salt reactor. The
proposed design temperature of the coolant at the
reactor exit is 10001C. The graphite blocks are
compatible with uoride salts as coolant. The reactor
concept is designed for atmospheric pressure opera-
tion. This design uses Ni-based high-temperature
alloys that have been similarly adopted for molten
salts. The reactor is proposed to be built in large sizes
(2000 MW
th
) with passive safety systems for decay heat
removal.
2.4. Modular helium reactor (MHR)
The MHR is a thermal reactor that can be used both
for hydrogen and electricity production in modules of
600 MW
th
. Its core consists of prismatic blocks of
graphite that allow coolant ow and contains ceramic
fuel. The temperature of the coolant, essentially He gas,
at the reactor exit is currently designed to achieve
temperatures around 8501C. It is proposed to achieve
10001C in the future within a new design, based on the
same reactor concepts, called the very high-temperature
reactor. The operating pressure of the MHR is 7 MPa.
The core design can provide passive safety by achieving
high temperatures during transients and by large
thermal inertia.
3. NUCLEAR-BASED HYDROGEN
PRODUCTION METHODS
To produce hydrogen, the hydrogen bonds in hydro-
carbons or water must be broken and hydrogen must
be separated from the reaction mixture. The most
efcient method for meeting increasing energy needs
could be to convert nuclear power into electricity and
hydrogen, thus providing effective and universal
energy carriers. Nuclear power plants produce heat
that can be used directly or converted to electricity for
the production of hydrogen. Four classes of H
2
production options are under development (e.g. [79]):
electrolysis:
conventional electrolysis (electricity1H
2
O
[liquid]-H
2
1O
2
)
high-temperature electrolysis (electricity1H
2
O
[steam]-H
2
1O
2
)
hybrid cycles (electricity1heat1H
2
O-[cyclic
chemical reactions]-H
2
1O
2
), and
thermochemical cycles (heat1H
2
O-[cyclic
chemical reactions]-H
2
1O
2
).
The near-term option is electrolysis. The long-term
options involve using heat to convert water to hydro-
gen and oxygen. Because heat is less expensive than
electricity (because the cost is avoided of converting
heat to electricity with associated losses), these ad-
vanced processes have the long-term potential of lower
production costs.
Estimates from Japan are that the cost of nuclear
thermochemical H
2
production could be as low as 60%
of that for nuclear H
2
production by the electrolysis of
water [7]. At the most fundamental level, thermo-
chemical H
2
production involves the conversion of
thermal energy to chemical energy (H
2
), whereas elec-
trolysis involves the conversion of thermal energy to
electricity and subsequent conversion of electricity to
chemical energy.
Nuclear energy provides a source of heat to produce
H
2
. Multiple processes are being investigated to pro-
duce H
2
from water and heat. If nuclear energy is to be
used for H
2
production, the nuclear reactor must de-
liver the heat at conditions that match the require-
ments imposed by the H
2
production process. At this
stage of development, it is unclear which chemical
processes will be the most economic; thus, the major
candidate technologies were examined to determine if
they impose similar requirements on the reactor [9].
Methods for obtaining hydrogen using carbon com-
pounds as the raw material will probably be the main
methods in the near future. However, the raw materials
and ecological limitations of steam conversion of
methane are stimulating the development of processes to
produce hydrogen from water. The most interesting of
these methods in the context of nuclear power are elec-
trolysis and thermochemical and thermoelectrochemical
cycles. Figure 1 show a brief overview of nuclear-based
hydrogen production technologies. The main processes
for hydrogen production include steam reforming of
natural gas, catalytic decomposition of natural gas,
partial oxidation of heavy oil, coal gasication, water
electrolysis, thermochemical cycles, and photo-chemical,
electrochemical and biological processes. The rst four
processes are based on fossil fuels.
In this paper, the main focus is given to the potential
thermochemical cycles for nuclear-based hydrogen
production as summarized below [2730].
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
125
3.1. Thermochemical and hybrid cycles
Several review articles and reports on thermochemical
hydrogen production have been published [1420],
including a comprehensive listing of past and present
activities [31]. Specic categories of thermochemical
production of hydrogen have also been reviewed,
including solar [2325] and nuclear [32].
Thermochemical processes include a thermal dis-
association of water into hydrogen and oxygen through
a series of thermally driven chemical reactions. Overall
reaction becomes
H
2
O H
2
1
1
2
O
2
(1)
The purpose is to produce hydrogen at lower tem-
peratures than that for thermolysis of water, which takes
place at temperatures greater than 250030001C [33].
Figure 2 illustrates the reaction temperature levels
(from 100 to 30001C, respectively) of thermochemical
cycles which are potentially driven by various heat
source options as some examples. This gure shows
that the number of reaction steps decreases with an
increase in the temperature. The number of reaction
steps also indicates which cycles may be coupled with
solar, nuclear or geothermal options. It is also clear
that the higher the temperature level the lower reduce
the number of steps in a cycle. One thing is also clear
that thermochemical cycles offer a great opportunity to
realize hydrogen economy in a more efcient and more
environmentally benign manner.
In the literature, hundreds of individual cycles
have been proposed for hydrogen production during
the past 40 years. At present, the most promising
high-temperature cycles appear to be the sulfuriodine
(S-I) cycle and the BrCaFe cycle. Also, one of the
most important low-temperature cycles is copper
chlorine (CuCl) [e.g. [34]]. All these cycles were exa-
mined and analyzed in terms of their thermodynamic
feasibilities. In a recent study, only 25 cycles were
found feasible by Brown et al. [13]. The specic aspects
of thermochemical water decomposition processes
have also been investigated recently. For the sulfur
iodine (S-I) cycle utilizing all uid reagents, a higher
temperature (8259001C) is employed for the oxygen-
evolving reaction, while higher efciencies are possible.
The thermochemical SI cycle has been studied in
more detail elsewhere [35,36]. This cycle has in fact
been fully demonstrated in both Japan and the United
States and has been shown to be technically viable.
However, the commercial viability of any of these
cycles has yet to be demonstrated [37].
The copperchlorine (CuCl) cycle has recently been
proven in the laboratory, and several commercially
appealing variants are being evaluated. Recently,
Atomic Energy of Canada Limited (AECL) and the
Argonne National Laboratory in the United States
have developed a low-temperature cycle, to accom-
modate heat sources around 3505501C. For this
temperature range, the CuCl cycle appears to be one
of the most promising. This cycle has an estimated
efciency of 4060% at an envisioned operating
temperature of 3505501C. Development of the low-
temperature CuCl cycle was proposed during this
early research period. Although numerous studies have
been conducted on CuCl thermochemical cycle, only
a few recent ones are given in References [2730,34,38].
Some landmark type studies on energy, exergy, and
cost analyses of CuCl thermochemical cycle driven
by nuclear process heat were performed by Orhan
et al. [2729].
Research on thermochemical hydrogen production
is undertaken in many countries [3942]. Rosen [39]
reported that technologies for thermochemical hydrogen
production are being investigated by several countries
(Canada, Japan, United States, and France). In Canada,
technologies for thermochemical hydrogen production
are being investigated by AECL that couple with nuclear
reactor. Sandia National Laboratory in the United
States and the Atomic Energy Commission in France
Nuclear-based
hydrogen
production
options
Electrolysis
Thermochemical
Processes
Hybrid Processes
Figure 1. Some potential options of nuclear-based hydrogen
production.
Figure 2. Number of thermochemical reaction steps, depending on cycle temperature.
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
126 Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
are developing a hydrogen pilot plant with a sulphur-
iodine (S-I) cycle. The Korea Atomic Energy Research
Institute (KAERI) is collaborating with China to pro-
duce hydrogen with the 10-MW HTGR (HTR-10). The
Japan Atomic Energy Agency (JAEA) plans to complete
a large S-I plant to produce 60 000 m
3
h
1
of hydrogen
by 2020, an amount sufcient for about 1 million fuel
cell vehicles. Research in the United States on processes
for hydrogen production at temperatures below 5501C,
including thermochemical cycles, has been reported re-
cently [33]. Balta et al. [30] investigated various geo-
thermal-based hydrogen production methods, and they
dened four potential methods including hydrogen
production through thermochemical cycles.
In this study, we also investigate some potential low-
temperature thermochemical and hybrid cycles for
nuclear-based hydrogen production at temperatures
below 5501C, due to their wider practicality, and ex-
amine them as a sustainable option for hydrogen
production using nuclear resources. Zamrescu and
Dincer [17] and Granowskii et al. [18] have studied the
use of mechanical compression heat pumps and che-
mical heat pumps as one of the potential options to
upgrade the heat to reach higher temperatures for such
thermochemical cycles.
Such low-temperature thermochemical and hybrid
cycles, which can be coupled with process or waste heat
source, are listed in more detail in Table I. The simple
concepts of low-temperature cycles are illustrated in
Figure 3(ad), respectively, which are drawn based on
some literature data and information, as reported
elsewhere [30,41,4346].
3.1.1. The copperchlorine cycle. As mentioned, the
copperchlorine (CuCl) cycle was originally proposed
in the 1970s and has recently been proven at laboratory
level. The CuCl cycle is thus considered the most
promising low-temperature cycle and offers a number
of potential advantages over other cycles, such as [45]:
The maximum cycle temperature (5001C) allows
the use of a wider range of heat sources such as
geothermal, nuclear, or solar.
The recycling chemicals are relatively safe,
inexpensive, and abundant.
All reactions have been proven in the laboratory
and no signicant side reactions have been
observed.
A simple conceptual drawing of the CuCl cycle is
shown in Figure 3(a). It basically consists of ve main
steps:
(a) HCl
(g)
production step with equipment like a
uidized bed,
(b) O
2
production step,
Table I. Low-temperature thermochemical and hybrid cycles.
Cycle Name
Number of
steps Name of steps Reactions
Reaction
temperatures (1C) References
1 Copperchlorine 5 HCl production 2CuCl
2
1H
2
O-CuOCuCl
2
12HCl 400 [41]
O
2
production CuOCuCl
2
-2CuCl11/2O
2
500
Cu production 4CuCl1H
2
O-2CuCl
2
12Cu 2580
Drying CuCl
2
(aq)-CuCl
2
(s) 4100
H
2
production 2Cu12HCl-2CuCl1H
2
430475
2 Lithiumnitrite 3 HI production LiNO
2
1I
2
1H
2
O-LiNO
3
12HI 25 [43]
H
2
production 2HI-I
2
1H
2
425
O
2
production LiNO
3
-LiNO
2
11/2O
2
475
3 Magnesium
Chloride
3 HCl production MgCl
2
1H
2
O-2HCl1MgO 450 [44]
O
2
production MgO1Cl
2
-MgCl
2
11/2O
2
500
H
2
production 2HCl-Cl
2
1H
2
80
4 Heavy-element
halide
4 Decomposition of
UO
2
Br
2
3H
2
O
2(UO
2
Br
2
3H
2
O)-2UO
3
H
2
O1
4HBr12H
2
O
300 [45]
H
2
production 4EuBr
2
14HBr-4EuBr
3
12H
2
Exothermic
Decomposition of
EuBr
3
4EuBr
3
-4EuBr
2
12Br
2
300
O
2
production 2UO
3
H
2
O12Br
2
14H
2
O-
2(UO
2
Br
2
3H
2
O)1O
2
Exothermic
5 Sulphuric acid 3 H
2
production 2H
2
O1SO
2
-H
2
SO
4
1H
2
80 [45]
SO
3
production H
2
SO
4
-H
2
O1SO
3
450
O
2
production SO
3
-SO
2
11/2O
2
550
6 Cesium (Astrojet) 4 H
2
production 2H
2
O12Cs-2CsOH1H
2
450 [46]
CsO
2
production 2CsOH13/2O
2
-2CsO
2
1H
2
O 250
O
2
production 2CsO
2
-Cs
2
O13/2O
2
450
O
2
and Cs production Cs
2
O-2Cs11/2O
2

Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
127
(c) Cu production step,
(d) drying step, and nally
(e) hydrogen production step.
These chemical reactions form a closed internal loop
that recycles all of the copperchlorine compounds on
a continuous basis, without emitting any greenhouse
gases externally to the atmosphere [41]. One of the
reactions in this CuCl cycle is electrochemical, which
affects a considerable amount of energy cost. However,
the electrolytic step requires voltages signicantly
lower than needed for direct water electrolysis. Petri
et al. [33] mentioned that copperchlorine cycle has
been examined at Argonne National Laboratory,
consisting of three thermal reactions and one electro-
lytic reaction [38]. Experimental work has been per-
formed at Argonne National Laboratory to study the
reaction kinetics for the hydrogen and oxygen pro-
duction reactions. The experiments were conducted in
beds of solid material with a continuous ow of excess
gaseous reactants. The individual steps in the CuCl
cycle have been conrmed, the kinetics of the hydrogen
and oxygen generation reactions have been studied,
and the temperatures of the reaction steps have been
veried (e.g. [33]).
3.1.2. The lithiumnitrite cycle. The Argonne National
Laboratory has initially proposed this cycle, which
consists of three main steps: (i) hydrogen iodide (HI)
production step, (ii) hydrogen production step, and
(iii) oxygen production step. A schematic representa-
tion of this cycle is given in Figure 3(b). As can be seen
in this gure, lithium nitrite is oxidized by iodine at
about 251C to produce lithium nitrate and HI, so that
HI produces through thermal dissociation and lithium
nitrate produces oxygen through thermal decomposi-
tion, respectively. The oxidation of lithium nitrite by
iodine involved in this cycle, however, can hardly be
described as proceeding smoothly [42]. The rst step
proceeds slowly and similar to the well-studied
oxidation of sulphurous acid by iodine to form the
sulphate ion and HI. There is no reason to believe the
reaction will not go as suggested by Abraham and
Schreiner [43]. The second step in the cycle, the thermal
decomposition of HI, is well known. At the tempera-
ture indicated 23% dissociation will occur. Separation
of HI from the mixture can easily be accomplished by
thermal means such as distillation [43]. The third step
of the reaction may possibly produce LiO
2,
which is
very corrosive compound, under certain reaction
conditions [47]. This cycle has not yet been developed
to the extent of commercial application yet. Further
details and references to the chemistry of the reactions
on this cycle can be found elsewhere (e.g. [48]).
3.1.3. The magnesiumchloride cycle. The magne-
siumchloride (Reverse Deacon Cycle) has been
developed by Simpson et al. [44] at Idaho National
Laboratory and Argonne National Laboratory. The
magnesium-chloride cycle basically consists of three
reactions: two thermochemical and one electrolytic. The
concept of this cycle is illustrated in Figure 3(c). The
thermochemical reactions sum to the reverse Deacon
reaction. The electrolytic step involves the electrolysis of
Figure 3. Illustrations of four potential cycles (see details in Reference [30]).
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
128 Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
anhydrous HCl instead of aqueous HCl. Motupally
et al. [49] specied that anhydrous HCl electrolysis
requires signicantly less electrical energy than aqueous
HCl electrolysis. This cycle steps are dened by Petri
et al. [33]. They mentioned that HCl was successfully
generated through step 1, while step 2 has not been
tested, but is thermodynamically possible. Higher
temperatures, though, may be required to eliminate
undesirable side steps. Step 3 has been demonstrated
and optimized by Weidners team at the University of
South Carolina. Further proof-of-principle tests must be
run to establish the viability of this process. Further-
more, Simpson et al. [44] analyzed preliminarily the
efciency aspects. The proof-of-concept and lab-scale
experiments have been performed. Based on their results
and temperature range of the reactions, one may
indicate that the cycle is a promising route to low-
temperature hydrogen production although there is no
any ongoing work in this regard.
3.1.4. The hybrid sulphur (sulphuric acid) cycle. The
hybrid sulphur cycle (sulphuric acid cycle) has been
developed by the JAEA. This cycle mainly consists of
three steps: (i) hydrogen production step, (ii) sulphur
trioxide (SO
3
) production step and (iii) oxygen produc-
tion step. The steps 1 and 3 are electrolytic reactions.
Schematic concept of this cycle can be seen in Figure 3(d).
The system is based on a sulphuric acid (H
2
SO
4
) process,
which was developed as the Westinghouse process. The
sulphur trioxide (SO
3
) decomposition process is facilitated
by electrolysis using a solid electrolyte that conducts
oxygen ions. By this way, the temperature of the step can
be reduced approximately to 2501C compared with the
Westinghouse process. The details about the cycle may be
found elsewhere [45]. The theoretical thermal efciency of
the system based on chemical reactions depending on the
H
2
SO
4
concentration and heat recovery was in the range
of 3555%. In the Argonne National Laboratory,
improved SO
3
electrolysis cells have been developed to
lower the needed voltage and increase overall efciency
for step 3. There is a remarkable potential to produce
hydrogen using this cycle as the cycle operates at low
temperature.
3.1.5. The U-Eu-Br heavy-element halide cycle. The
cycle is based on heavy-element halide chemistry with a
maximum reaction temperature of 3001C. Heavy-
element halide cycle is the lowest known temperature
for a purely thermochemical hydrogen production.
The heavy-element halide cycle essentially consists of
four steps. The steps 1 and 3 are endothermic, whereas
steps 2 and 4 are exothermic. This cycle was performed
at Argonne National Laboratory to determine the each
step (e.g. [45]).
Step 1: Determine the chemical products
that result from thermal decomposition of
UO
2
Br
2
3H
2
O.
Step 2: Investigate and model the factors that
inuence the reaction of Eu
21
ions with H
1
ions
in aqueous hydrobromic acid to generate H
2
gas.
Step 3: Study the thermal reduction of EuBr
3
to
EuBr
2
and establish the degree of completion
at 3001C and whether a potentially interfering
EuOBr impurity is produced;
Step 4: Determine the chemical consequences of
reacting hydrated uranium trioxide (UO
3
H
2
O
(s)) with an excess amount of bromine water (Br
2
dissolved in H
2
O).
As a result, thermodynamic data of this system
are not well known. An engineering application of
the heavy-element halide cycle needs to improve for
corrosiveness of the chemicals. The low operating
temperature of 3001C cycle can be more applicable
to nuclear-based hydrogen production systems, rather
than for higher temperature cycles.
3.1.6. The astrojet central cycle. One of the potential
methods for thermochemical hydrogen production
involves the use of cesium through the astrojet central
cycle [46]. This thermochemical cycle consists of four
steps with a maximum reaction temperature of 4501C.
It basically consists of four chemical reactions as their
steps, namely;
(a) hydrogen production,
(b) CsO
2
production,
(c) oxygen production, and
(d) oxygen and cesium production.
This cycle is technically a closed cycle, in which
water is thermally decomposed and all the other
chemicals are recycled. The ideal hydrogen production
efciency for this cycle was estimated as 66% [50],
where the efciency was dened as the higher heating
value (HHV) of combustion of H
2
(-286 kJ mol
1
)
divided by the external thermal heat supplied to the
process per gram mole of H
2
produced. A practical
efciency of 45% was estimated by [50] for the process
based on the thermal dissociation of Cs
2
O. It is
important to note that if dissociation and separation
were accomplished by the high-temperature electrolysis
of molten Cs
2
O, the efciency would be calculated to
be 31.6% [50].
4. THERMODYNAMIC ANALYSIS
Efcient use of energy is a signicant contributor to any
sustainable plan for meeting growing energy demands.
Therefore, it is essential to evaluate the technologies for
nuclear-based hydrogen production in terms of energy
and exergy efciencies. The overall efciencies of the
cycles studied for utilizing nuclear process/water heat for
this purpose depend on parameters such as operating
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
129
temperature, recycling ratio and heat loss. We thermo-
dynamically analyze four low-temperature cycles in this
study for comparison purposes.
In the thermodynamic analysis, we consider 1 kmol
of hydrogen produced per each cycle considered;
hence, all quantities are provided in terms of per kmol
of hydrogen produced. In addition, we assume the
following:
The values for the reference environment (dead
state) temperature (T
0
) and pressure (P
0
) are 251C
and 1 atm, respectively.
In all chemical reactions, reactants and products
are at the reaction temperature and a pressure of
1 atm.
All processes are considered steady-state and
steady-ow with negligible potential and kinetic
energy effects.
The heat loss through the cycle is assumed as 30%.
For a general steady-state, steady-ow process, the
three balance equations, namely mass, energy, and
exergy balance equations, are employed to nd the
work input, the rate of exergy destruction, energy and
energy efciencies.
As mass is conserved in chemical reactions, the mass
of products and reactants are equal and in general, the
mass balance equation can be expressed in the rate
form as
X
_ m
in
=
X
_ m
out
or
X
_ m
R
=
X
_ m
P
(2)
where _ m is the mass ow rate, and the subscript in
stands for inlet and out for outlet.
For a steady-state process, the general energy
balance can be expressed below as the total energy
input equal to total energy output:
X
_
E
in
=
X
_
E
out
(3)
The heat transfer for a chemical process involving
no work interaction W is determined from the energy
balance
P
_
E
in
=
P
_
E
out
applied to the system with
W50. For a steady-state reaction process, the energy
balance reduces to
Q =
X
n
P
(

h
o
f
1

h

h
o
)
P

X
n
R
(

h
o
f
1

h

h
o
)
R
(4)
and the exergy balance for the steady-state process,
involving chemical reactions, becomes
X
_
Ex
in
=
X
_
Ex
out
(5)
where
X
_
Ex
out
=
_
Ex
H2
1
_
Ex
d
1
_
Ex
l
(6)
The exergy associated with a process at specied
state is the sum of two contributions: physical and
chemical. Thus, the specic exergy of the process is
calculated by
ex = (h h
0
) T
0
(s s
0
)1ex
ch
(7)
where h is enthalpy, s is entropy, and the subscript zero
indicates properties at the reference (dead) state of P
0
and T
0
.
The exergy rate is calculated by
_
Ex = _ mex (8)
Combining Equations (6) and (7) yields
ex
d
=
X
[(h h
0
) T
0
(s s
0
)1ex
ch
]
in

X
[(h h
0
) T
0
(s s
0
)1ex
ch
]
out
1 1
T
0
T
r

Q (9)
After writing mass, energy, and exergy balances
for the system, enthalpy values of compounds are
evaluated using the Shomate equations [51] as follows:

h

h
0
=A T1B
T
2
2
1C
T
3
3
1D
T
4
4
E
1
T
1F H (10)
s =A ln(T)1B T1C
T
2
2
1D
T
3
3
E
1
2T
2
1G (11)
where T is 1/1000 of the specied temperature (in K)
of compound and A, B, C, D, E, F,G, and H are
constants, as given in Table II for each compound.
With the specic enthalpy and entropy values, we
can calculate the specic chemical exergy ex
ch
value of
each compound. The chemical exergy based on a
typical exergy reference environment exhibiting stan-
dard values of the environmental temperature T
0
and
pressure P
0
such as 251C and 1 atm is called standard
chemical exergy. The values of the standard chemical
exergies for the reactants and products are taken from
the literature [52], as listed in Table III. However, there
is no information about LiNO
2
and LiNO
3
in this
Ref. [52]; therefore, these data are assumed as constant
and taken from [53].
To determine the standard chemical exergy of any
substance not present in the environment, we consider
the reaction of the substance with other substances for
which the standard chemical exergies are known, and
write
ex
ch
= DG1
X
P
nex
ch

X
R
nex
ch
(12)
where DG is the change in Gibbs function for the
reaction, regarding each substance as separate at
temperature T
0
and pressure P
0
. The other two terms
on the right side of Equation (12) are evaluated using
the known standard chemical exergies, together with
values of n, which denotes the moles of these reactants
and products per mole of the substance whose chemi-
cal exergy is being evaluated.
The energy efciency of the overall cycle can be
dened as the total energy output of hydrogen to the
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
130 Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
T
a
b
l
e
I
I
.
E
n
t
h
a
l
p
y
o
f
f
o
r
m
a
t
i
o
n
,
r
e
f
e
r
e
n
c
e
e
n
t
r
o
p
y
,
a
n
d
S
h
o
m
a
t
e
c
o
n
s
t
a
n
t
s
f
o
r
c
h
e
m
i
c
a
l
c
o
m
p
o
u
n
d
s
.
C
o
m
p
o
u
n
d

h
o f
(
k
J
k
m
o
l

1
)
s
0
(
k
J
k
m
o
l

1
K

1
)
A
B
C
D
E
F
G
H
C
l
2
(
g
)
0
2
2
3
.
0
8
3
3
.
0
5
0
6
1
2
.
2
2
9
4

1
2
.
0
6
5
1
4
.
3
8
5
3
3

0
.
1
5
9
4
9
4

1
0
.
8
3
4
8
2
5
9
.
0
2
9
0
C
u
(
s
)
0
3
3
.
1
7
1
7
.
7
2
8
9
1
2
8
.
0
9
8
7

3
1
.
2
5
2
8
9
1
3
.
9
7
2
4
3
0
.
0
6
8
6
1
1

6
.
0
5
6
5
9
1
4
7
.
8
9
5
9
2
0
H
2
(
g
)
0
1
3
0
.
6
8
3
3
.
0
6
6
1

1
1
.
3
6
3
4
1
1
.
4
3
2
8
1
6

2
.
7
7
2
8
7
4

0
.
1
5
8
5
5
8

9
.
9
8
0
7
9
7
1
7
2
.
7
0
7
9
0
O
2
(
g
)
0
2
0
5
.
0
7
2
9
.
6
5
9
6
.
1
3
7
2
6
1

1
.
1
8
6
5
2
1
0
.
0
9
5
7
8

0
.
2
1
9
6
6
3

9
.
8
6
1
3
9
1
2
3
7
.
9
4
8
0
I
2
(
g
)
6
2
4
2
0
2
6
0
.
6
9
3
7
.
7
9
7
6
3
0
.
2
2
5
4
5
3

0
.
9
1
2
5
5
6
1
.
0
3
4
9
1
3

0
.
0
8
3
8
2
6
5
0
.
8
6
8
6
5
3
0
5
.
9
1
9
9
6
2
.
4
2
1
1
I
2
(
l
)
1
3
5
2
0
1
5
0
.
3
6
8
0
.
6
6
9
1
9
6
.
8
5
5
7
E

0
8

8
.
7
2
4
E

0
8
3
.
7
2
3
1
E

0
8
4
.
7
3
5
8
3
E

1
0

1
0
.
5
2
7
8
2
2
4
7
.
9
7
9
8
1
3
.
5
2
3
0
2
H
I
(
g
)
2
6
3
6
0
2
0
6
.
5
9
2
6
.
0
4
5
4
4
.
6
8
9
6
7
8
4
.
9
1
1
7
6
5

2
.
6
5
4
3
9
7
0
.
1
2
1
4
1
9
1
8
.
7
5
4
9
9
2
3
7
.
2
0
1
8
2
6
.
3
5
9
0
3
H
2
O
(
l
)

2
8
5
8
3
0
6
9
.
9
5

2
0
3
.
6
0
6
1
5
2
3
.
2
9

3
1
9
6
.
4
1
3
2
4
7
4
.
4
5
5
3
.
8
5
5
3
2
6

2
5
6
.
5
4
7
8

4
8
8
.
7
1
6
3

2
8
5
.
8
3
0
4
H
2
O
(
g
)

2
4
1
8
3
0
1
8
8
.
8
4
3
0
.
0
9
2
6
.
8
3
2
5
1
4
6
.
7
9
3
4
3
5

2
.
5
3
4
4
8
0
.
0
8
2
1
3
9

2
5
0
.
8
8
1
2
2
3
.
3
9
6
7

2
4
1
.
8
2
6
4
C
u
C
l
2
(
s
)

2
0
5
8
5
0
1
0
8
.
0
6
7
0
.
2
1
8
8
2
2
3
.
3
6
1
3
2

1
4
.
8
6
8
7
6
4
.
0
5
3
8
9
9

0
.
3
6
6
2
0
3

2
2
8
.
9
4
0
5
1
8
4
.
6
3
7
8

2
0
5
.
8
5
3
2
C
u
O
(
s
)

1
5
6
0
6
0
4
2
.
5
9
4
8
.
5
6
4
9
4
7
.
4
9
8
6
0
7

0
.
0
5
5
9
8
0
.
0
1
3
8
5
1

0
.
7
6
0
0
8
2

1
7
3
.
4
2
7
2
9
4
.
8
5
1
2
8

1
5
6
.
0
6
3
2
C
u
C
l
(
l
)

1
3
1
1
8
0
9
3
.
7
5
6
6
.
9
4
4

3
.
6
9
E

1
0
2
.
1
6
E

1
0

3
.
9
0
E

1
1

9
.
8
1
E

1
2

1
5
1
.
1
3
7
4
1
7
4
.
7
6
5
3

1
3
1
.
1
7
8
C
u
C
l
(
s
)

1
3
8
0
7
0
8
7
.
0
4
7
5
.
2
7
1

2
6
.
8
3
2
1
2
2
5
.
6
9
1
5
6

7
.
3
5
7
9
8
2

1
.
8
4
7
7
4
7

1
6
5
.
7
2
9
9
1
7
4
.
6
6
4
4

1
3
8
.
0
7
2
H
C
l
(
g
)

9
2
3
1
0
1
8
6
.
9
3
2
.
1
2
3
9
2

1
3
.
4
5
8
0
5
1
9
.
8
6
8
5
2

6
.
8
5
3
9
3
6

0
.
0
4
9
6
7
2

1
0
1
.
6
2
0
6
2
2
8
.
6
8
6
6

9
2
.
3
1
2
0
1
M
g
O
(
s
)

6
0
1
2
4
0
2
6
.
8
5
4
7
.
2
5
9
9
5
5
.
6
8
1
6
2
1

0
.
8
7
2
6
6
5
0
.
1
0
4
3

1
.
0
5
3
9
5
5

6
1
9
.
1
3
1
6
7
6
.
4
6
1
7
6

6
0
1
.
2
4
0
8
M
g
C
l
2
(
s
)

6
4
1
6
2
0
8
9
.
6
2
7
8
.
3
0
7
3
3
2
.
4
3
5
8
8
8
6
.
8
5
8
8
7
3

1
.
7
2
8
9
6
7

0
.
7
2
9
9
1
1

6
6
7
.
5
8
2
3
1
7
9
.
2
6
3
9

6
4
1
.
6
1
6
4
H
2
S
O
4
(
g
)

7
3
5
1
3
0
2
9
8
.
7
8
4
7
.
2
8
9
2
4
1
9
0
.
3
3
1
4

1
4
8
.
1
2
9
9
4
3
.
8
6
6
3
1

0
.
7
4
0
0
1
6

7
5
8
.
9
5
2
5
3
0
1
.
2
9
6
1

7
3
5
.
1
2
8
8
S
O
3
(
g
)

3
9
5
7
7
0
2
5
6
.
7
7
2
4
.
0
2
5
0
3
1
1
9
.
4
6
0
7

9
4
.
3
8
6
8
6
2
6
.
9
6
2
3
7

0
.
1
1
7
5
1
7

4
0
7
.
8
5
2
6
2
5
3
.
5
1
8
6

3
9
5
.
7
6
5
4
S
O
2
(
g
)

2
9
6
8
4
0
2
4
8
.
2
1
2
1
.
4
3
0
4
9
7
4
.
3
5
0
9
4

5
7
.
7
5
2
1
7
1
6
.
3
5
5
3
4
0
.
0
8
6
7
3
1

3
0
5
.
7
6
8
8
2
5
4
.
8
8
7
2

2
9
6
.
8
4
2
2
S
o
u
r
c
e
:
A
d
a
p
t
e
d
f
r
o
m
[
5
1
]
.
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
131
total energy input by
Z
overall
=
(1 r) _ m
H2
LHV
H2
_
W
e;in
1
P
_
Q
in
(13)
where LHV
H2
is the lower heating value per kmol of
hydrogen and r is the total recycling ratio of the
chemical reaction, dened by the number of kmols of
unreacted substances over the number of kmols of
supplied reactants. The lower heating value of hydro-
gen is taken as 239.92 MJ kmol
1
H
2
.
An exergy balance can be used in formulating exergy
efciency for the cycles. At steady state, the rate at
which exergy enters the cycles equals to the rate at
which exergy exits plus the rate at which exergy is
destroyed within the system. We assume a heat loss of
30% occurred.
The exergy efciency of the overall cycle is dened,
as the ratio of the exergy output of the hydrogen (H
2
)
to the total exergy input required for the cycle, as
follows:
c
overall
=
(1 r)
_
Ex
H2
_
W
e;in
1
P
_
Ex
in
(14)
Here, we also investigate some parameters about
sustainability and system improvement potential.
These are expected to give a comparative perspective
about the process and application and help people to
see how they will contribute to the environment and
sustainable development.
In this regard, sustainable development requires not
only that the sustainable supply of clean and afford-
able energy resources be used but also the resources
should be used efciently. Exergy methods are very
useful tools for improving efciency, which maximize
the benets and usage of resources and also minimize
the undesired effects (such as environmental damage).
Exergy analysis can be used to improve the efciency
and sustainability [54]. Rosen et al. [55] dened a
relation between exergy efciency (c) and the sus-
tainability index (SI) as
c = 1
1
SI
(15)
SI =
1
1 c
(15a)
which shows how sustainability increases with the
exergy efciency of a process increases.
Furthermore, Van Gool [56] proposed that maxi-
mum improvement in the exergy efciency for a pro-
cess or system is obviously achieved when the exergy
loss or irreversibility rates (
_
Ex
in

_
Ex
out
) are mini-
mized. Consequently, he suggested that it is useful
to employ the concept of an exergetic improvement
potential when analyzing different processes or sectors
of the economy. This improvement potential in the rate
form, denoted by I
_
P, is given by
I
_
P = (1 c)(
_
Ex
in

_
Ex
out
) (16)
5. RESULTS AND DISCUSSION
In this paper, some signicant low-temperature thermo-
chemical and hybrid cycles, are listed in Table I, are
reviewed and examined for potential applications with
nuclear thermal energy and hence analyzed thermo-
dynamically through energy and exergy. The parametric
studies with variable recycling ratios and heat losses
are performed, and some critical results are given in
Figures 4 and 5. These cycles appear to play some
crucial role in getting integrated with nuclear-based
process and waste heat for hydrogen production, which
is particularly aimed to help hydrogen economy. Using
energy and exergy balances, both energy and exergy
efciencies of the process cycles are calculated. Figure 4
shows the energy (a) and exergy (b) efciencies asso-
ciated with various recycling ratios from 0 to 0.9.
An important parameter is recycling ratio that
affects the cycles energy and exergy efciencies.
However, there is no enough information about the
recommended recycling ratios of the analyzed cycles in
the open literature. In this context, we performed a
parametric study to investigate the cycle performance
for a range of practical recycling ratio. Of course, these
will provide an application possibility for future.
Therefore, in this study, we investigate how the
energy and exergy efciencies in the studied cycles
vary with changing recycling ratio. The heat ow for
the cycles per kmol of hydrogen is calculated from
Equation (5). The energy efciencies of the each
cycle are determined to about 50% based on complete
reaction and lower heating value (LHV) using
Equation (13). In this calculation, the auxiliary works
are not considered.
Table III. Gibbs free energy of formation and standard chemical
energy for each compound.
Compounds
Specic Gibbs free energy
of formation (kJ kmol
1
)
Standard chemical
exergy (kJ kmol
1
)
H
2
O (g) 228638 9437
H
2
O (l) 237170 920
CuCl
2
(s) 161689 94 510
HCl (g) 95296 84 549
CuO(s) 128312 6273
CuCl (l) 115994 78 406
CuCl (s) 120884 73 516
MgO (s) 568950 62 405
MgCl
2
(s) 592113 160 857
H
2
SO
4
(g) 691375 197 630
SO
3
(g) 371069 241 936
SO
2
(g) 300144 310 876
HI (g) 37453 242 868
Source: Reference [52].
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
132 Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
If we continue elaboration on the results, we can
indicate that, as introduced above, Figure 4 shows the
inuence of total recycling ratio on energy (a) and
exergy (b) efciencies, respectively. The following
cycles are included in the analysis presented in the
gure: CuCl, LiNO
2
, Mg-Cl and H
2
SO
4
. The highest
energy efciency, such as 51%, is obtained by CuCl
and H
2
SO
4
cycles. However, the exergy efciency of
CuCl cycle is better than that of H
2
SO
4
cycle with
65 vs 57%. Also, it can be seen in Figure 4, the cycle
exergy efciencies are slightly higher than their corres-
ponding energy efciencies, which are calculated based
on LHV
H2
. This difference may be caused by the effect
of inlet heat in Equation (13) and inlet exergy in
Equation (14). In these two equations, LHV of hydro-
gen is very close to its exergy concept as inlet exergy
includes also the chemical exergy of compounds, and
also inlet works are the same in both equations.
However, inlet heat in Equation (13) is higher than
inlet exergy in Equation (14) as inlet exergy includes
also the exergetic temperature factor. Thus, exergy
efciency is higher than energy efciency.
Figure 5 illustrates the effects of the heat loss on the
overall CuCl cycle energy (a) and exergy (b) ef-
ciencies. The energy efciency values for the CuCl
cycle vary between 4 and 60%, whereas the corres-
ponding exergy efciency values for the same tem-
perature range from 5 to 77%, respectively. Also, it is
clear in this gure that both energy and exergy ef-
ciencies decrease with the recycling ratio from 0 to 0.9.
As can be seen in Figure 4, for the CuCl cycle the
energy (a) and exergy (b) efciency values vary bet-
ween 551% and 665%, respectively, assuming 30%
of input energy as heat loss from the cycle while the
recycling ratio increases from 0 to 0.9. It is important
to note that the CuCl cycle efciency values are in a
good agreement with those reported by Rosen and
Scott [57] and Yildiz and Kazimi [58]. In the present
study, the energy efciency of thermochemical hydro-
gen production is found to be less than that of the
0
10
20
30
40
50
60
70
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
E
n
e
r
g
y

E
f
f
i
c
i
e
n
y

(
%
)
Recycling Ratio
0.1 Qloss
0.2 Qloss
0.3 Qloss
0.4 Qloss
0.5 Qloss
0
10
20
30
40
50
60
70
80
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
E
x
e
r
g
y

E
f
f
i
c
i
e
n
c
y

(
%
)
Recycling Ratio
0.1 Qloss
0.2 Qloss
0.3 Qloss
0.4 Qloss
0.5 Qloss
(a) (b)
Figure 5. Variation of energy (a) and exergy (b) efciencies with recycling ratio for the CuCl cycle.
Figure 4. Energy (a) and exergy (b) efciencies of cycles versus recycling ratio.
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
133
SMR technology [58]. But nuclear-based hydrogen
production using thermochemical cycles appears to be
a promising solution for the future hydrogen economy
to be run on renewable, and it will help to reduce
environmental impact (greenhouse gas emissions) and
increase sustainable development. Therefore, copper
chlorine cycle has been identied as a highly promising
option among all the cycles considered for thermo-
chemical hydrogen production.
Figure 6 shows the effects of recycling ratio on the
sustainability index of the four considered cycles. It is
clearly seen in this gure that the sustainability index
of the analyzed cycles decreases exponentially with the
recycling ratios. In this regard, at lower recycling
ratios, the corresponding sustainability of cycles
increases. The highest sustainability index, such as
2.85, is obtained by CuCl cycle, which is based on
complete reaction. The sustainability index for the
analyzed cycles varies from 1.0 to 2.85 with the
recycling ratios varying between 0 and 0.9.
Van Gools improvement potential on the rate
basis (I
_
P) given in Equation (16) is calculated for the
analyzed low-temperature thermochemical cycles and
given in Figure 7. It is found that the H
2
SO
4
cycle has
the highest I
_
P value with about 78 000 kW, followed by
LiNO
3
, MgCl and CuCl with about 60 500, 60 400
and 42 000 kW, respectively.
In summary, some key advantages about the CuCl
cycle can be listed as follows:
completed proof-of-principle experiments for all
of the steps,
maximum temperature requirement of the cycle
about 5001C,
thermodynamic data of this cycle which are well
known,
relatively safe, inexpensive, and abundant recycling
chemicals,
A relatively high-efciency operation (about
4050%) at an envisioned operating temperature
of 3505001C.
conceptual process design that uses commercially
practiced processes.
Other cycles, e.g. the lithiumnitrite cycle, the
magnesiumchloride cycle, the heavy-element halide
cycle and the sulphuric acid cycle have some value and
promise for future. Although no more information on
cycles 4 and 6 as listed in Table I is available, we still
feel that they have some potential for hydrogen pro-
duction using nuclear-based process/waste heat, but
require further studies and experiments.
5. CONCLUSIONS
In this paper, some potential thermochemical and
hybrid cycles for nuclear-based hydrogen production
have been discussed, with some key focus on their
operating conditions, temperature ranges, cycle pheno-
mena and performance aspects, and a brief perfor-
mance analysis through energy and exergy efciencies
has been presented for comparison purposes. Some
ndings may be summarized as follows:
The energetic and exergetic efciencies of the
CuCl cycle are obtained to be 51 and 65%,
respectively, based on the complete reactions.
The recycling ratio has an important effect on
the energy and exergy efciencies of the cycles.
As it increases, energy and exergy efciency values
decrease.
Low-temperature thermochemical and hybrid
cycles for hydrogen production, which can be
coupled with nuclear option, should be developed
and comprehensively analyzed in terms of their
thermodynamic feasibilities.
The sustainability index for the low-temperature
thermochemical cycles varies from 1.0 to 2.85
with the recycling ratios varying between 0 and
0
20000
40000
60000
80000
100000
120000
140000
0
I
m
p
r
o
v
e
m
e
n
t

P
o
t
e
n
t
i
a
l

R
a
t
e

(
k
W
)
Recycling Ratio
0.1 0.2 0.3 0.4 0.5
Figure 7. Variation of improvement potential with recycling
ratio for various thermochemical cycles.
0.0
0.5
1.0
1.5
2.0
2.5
3.0
3.5
0
S
u
s
t
a
i
n
a
b
i
l
i
t
y

I
n
d
e
x
Recycling Ratio
LiNO
CuCl
MgCl
H SO
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Figure 6. Variation of sustainability index with recycling ratio for
various thermochemical cycles.
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
134 Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
0.9. The highest sustainability index, such as 2.85,
is obtained by CuCl cycle, which is based on
complete reaction.
The CuCl cycle is one of the most pro-
mising ways to produce hydrogen efciently,
without emitting any greenhouse gases to the
atmosphere.
NOMENCLATURE
_
E 5energy rate (kW)
ex 5specic exergy (kJ kg
1
)
ex
ch
5specic chemical exergy (kJ kg
1
)
_
Ex 5exergy rate (kW)
G 5Gibbs function (kJ)
h 5specic enthalpy (kJ kg
1
)
h 5specic enthalpy (kJ kmol
1
)
h
o
5specic enthalpy at reference state
(kJ kmol
1
)
h
o
f
5specic enthalpy of formation
(kJ kmol
1
)
I
_
P 5improvement potential rate (kW)
LHV
H2
5lower heating value of H
2
(MJ kg
1
)
_ m 5mass ow rate (kg s
1
)
n 5number of moles (kmol)
Q 5heat (kJ)
r 5recycling ratio()
s 5specic entropy (kJ kg
1
K
1
)
SI 5sustainability index()
T 5temperature (K or 1C)
W 5work (kJ)
Greek symbols
Z 5energy efciency
c 5exergy efciency
DG 5change in Gibbs function for a
reaction (kJ)
Subscripts
d 5destruction
e 5electricity
in 5input, inlet
l 5loss
out 5output, outlet
P 5product
R 5reactant
s 5system
0 5reference or dead state
ACKNOWLEDGEMENTS
The author gratefully acknowledge the support pro-
vided by the Ontario Research Excellence Fund and
the Natural Sciences and Engineering Research Coun-
cil of Canada.
REFERENCES
1. Torjman M, Shaaban H. Nuclear energy as
a primary source for a clean hydrogen energy
system. Energy Conversion and Management 1998;
39:2732.
2. Verfondern K, Nishihara T. Safety aspects of the
combined HTTR/steam reforming complex for
nuclear hydrogen production. Progress in Nuclear
Energy 2005; 47:527534.
3. Forsberg C. Futures for hydrogen produced using
nuclear energy. Progress in Nuclear Energy 2005;
47:484495.
4. Ponomarev-Stepnoi NN. Nuclear-hydrogen power.
Atomic Energy 2004; 96:375385.
5. Onuki K, Inagaki Y, Hino R, Tachibana Y.
Research and development on nuclear hydrogen
production using HTGR at JAERI. Progress in
Nuclear Energy 2005; 47:496503.
6. Dincer I. Environmental and sustainability aspects
of hydrogen and fuel cell systems. International
Journal of Energy Research 2007; 31:2955.
7. Naterer GF, Fowler M, Cotton J, Gabriel K.
Synergistic roles of off-peak electrolysis and thermo-
chemical production of hydrogen from nuclear
energy in Canada. International Journal of Hydrogen
Energy 2008; 33:68496857.
8. Bertel E. Nuclear energythe hydrogen economy.
Nuclear Energy Agency News 2004; 22.
9. Duffey R, Miller A. From option to solution:
nuclear contribution to global sustainability using
the newest innovations. World Nuclear Association
Symposium, London, U.K., September 2002.
10. Duffey R. Green atoms. Power and Energy 2005;
2(2):812.
11. Dincer I. Technical, environmental and exergetic
aspects of hydrogen energy systems. International
Journal of Hydrogen Energy 2002; 27(3):265285.
12. Turner J, Sverdrup G, Mann MK, Maness PC,
Kroposki B, Ghirardi M, Evans RJ, Blake D.
Renewable hydrogen production. International
Journal of Energy Research 2008; 32:379407.
13. Brown LC, Funk JF, Showalter SK. Initial Screening
of Thermochemical Water-Splitting Cycles for High
Efciency Generation of Hydrogen Fuels Using
Nuclear Power: GAA23373, San Diego, CA, 2000.
14. Naterer GF, Gabriel K, Lu L. Recent advances
in nuclear based hydrogen production with the
thermochemical copper-chlorine cycle. Journal of
Engineering for Gas Turbines and Power-Transactions
of the ASME 2009; 131(3):032905.
15. Eisenstadt MM, Cox KE. Hydrogen production
from solar energy. Solar Energy 1975; 17:5965.
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
135
16. Costogue EN, Yasui RK. Performance data for
a terrestrial solar photovoltaic/water electrolysis
experiment. Solar Energy 1977; 19:205210.
17. Zamrescu C, Dincer I. Performance investigation
of high-temperature heat pumps with various BZT
working uids. Thermochimica Acta 2009; 488:
6677.
18. Granowskii M, Dincer I, Rosen MA, Pioro I.
Thermodynamic analysis of the use a chemical heat
pump to link a supercritical water-cooled nuclear
reactor and a thermochemical water-splitting cycle
for hydrogen production. Journal of Power and
Energy Systems 2008; 2:756767.
19. Maack MH, Skulason JB. Implementing the
hydrogen economy. Journal of Cleaner Production
2006; 14:5264.
20. McQuillan BW, Brown LC, Besenbruch GE,
Tolman R, Cramer T, Russ BE. High efciency
generation of hydrogen fuels using solar thermo-
chemical splitting of water. Annual Report,
GA-A24972, San Diego, CA, 2005.
21. Carty RH, Mazumder MM, Schreider JD,
Pangborn JB. Thermochemical Hydrogen Production.
Gas Research Institute for the Institute of Gas
Technology: Chicago, IL, 1981; 14. GRI-80/0023.
22. Yilanci A, Dincer I, Ozturk HK. A review on solar-
hydrogen/fuel cell hybrid energy systems for
stationary applications. Progress in Energy and
Combustion Science, DOI:10.1016/j.pecs.2008.07.004.
23. Steinfeld A. Solar thermochemical production
of hydrogena review. Solar Energy 2005; 78:
603615.
24. Fletcher EA. Solar thermal processing: a review.
Journal of Solar Energy Engineering 2001; 123:
6374.
25. Beghi GE. A decade of research on thermochemical
hydrogen at the Joint Research Centre, ISPRA.
International Journal of Hydrogen Energy 1986;
11(12):761771.
26. Forsberg CW, Peterson PF, Pickard PS. Molten
saltcooled advanced high temperature reactor for
production of hydrogen and electricity. Nuclear
Technology 2003; 144:289302.
27. Orhan MF, Dincer I, Rosen MA. The oxygen
production step of a copper-chlorine thermochemical
water decomposition cycle for hydrogen production:
energy and exergy analyses. Chemical Engineering
Science, DOI:10.1016/j.ces.2008.10.047.
28. Orhan MF, Dincer I, Rosen MA. Energy and exergy
assessments of the hydrogen production step of
a copperchlorine thermochemical water splitting
cycle driven by nuclear-based heat. International
Journal of Hydrogen Energy 2008; 33:64566466.
29. Orhan MF, Dincer I, Naterer GF. Cost analysis of
a thermochemical CuCl pilot plant for nuclear-
based hydrogen production. International Journal of
Hydrogen Energy 2008; 33:60066020.
30. Balta MT, Dincer I, Hepbasli A. Geothermal-
based hydrogen production using thermochemical
and hybrid cycles: a review and analysis. Inter-
national Journal of Energy Research 2010; 34(9):
757775.
31. Funk JE. Thermochemical hydrogen production:
past and present. International Journal of Hydrogen
Energy 2001; 26(3):185190.
32. Yalcin S. A review of nuclear hydrogen production.
International Journal of Hydrogen Energy 1989;
4(8):551561.
33. Petri MC, Yildiz B, Klickman AE. US work on
technical and economic aspects of electrolytic,
thermochemical, and hybrid processes for hydrogen
production at temperatures below 5501C. Interna-
tional Journal of Nuclear Hydrogen Production
Applications 2006; 1(1):7991.
34. Lewis MA, Masin JG, Vilim RB. Development of
the low temperature CuCl cycle. Proceedings of
2005 International Congress on Advances in Nuclear
Power Plants. ICAPP 005. American Nuclear
Society: Seoul, Korea, 1519 May 2005.
35. Forsberg CW. Hydrogen, nuclear energy and the
advanced high-temperature reactor. International
Journal of Hydrogen Energy 2003; 28:10731081.
36. Brown LC, Besenbruch GE, Lentsch RD, Schultz KR,
Funk JF, Pickard PS. High efciency generation of
hydrogen fuels using nuclear power: GA-A24285.
General Atomics Project 30047, 2003.
37. Basic research needs for the hydrogen economy.
Report of the Basic Energy Sciences Workshop
on Hydrogen Production, Storage and Use, 1315
May 2003. Available from: http://www.er.doe.gov/
bes/reports/les/NHE_rpt.pdf [accessed 16 September
2007].
38. Serban M, Lewis MA, Basco JK. Kinetic study for
the hydrogen and oxygen production reactions in the
copper chlorine thermochemical cycle. Conference
Proceedings, 2004 AIChE Spring National Meeting,
2004; 26902698.
39. Rosen MA. Exergy analysis of hydrogen production
by thermochemical water decomposition using the
Ispra Mark-10 Cycle. International Journal of
Hydrogen Energy 2008; 33:69216933.
40. Nuclear Production of Hydrogen. Proceedings of
the Third Information Exchange Meeting. OECD
Publishing: Oarai, Japan, 2006.
41. Naterer GF. Economics of a thermochemical pilot
plant for nuclear-produced hydrogen in Ontario,
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
136 Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
Technical Report, University of Ontario Institute of
Technology, Oshawa, Ontario, Canada, 2007.
42. Fujii KK, Kondo WK. Thermochemical process for
manufacture of hydrogen and oxygen from water.
U.S. Patent 3969493, 13 July 1976.
43. Abraham BM, Schreiner F. A low temperature
thermal process for the decomposition of water.
Science 1973; 180:959.
44. Simpson MF, Herrmann SD, Boyle BD. A hybrid
thermochemical electrolytic process for hydrogen
production based on the reverse deacon reaction.
International Journal of Hydrogen Energy 2006;
31:12411246.
45. International Atomic Energy Agency (IAEA).
Advanced applications of water cooled nuclear power
plants. IAEA-TECDOC-1584, Vienna, Austria, July
2008. ISBN 9789201058089.
46. Nakamura TM. Method and apparatus for generating
electricity magneto hydrodynamically U.S. Patent,
3980907, 14 September 1976.
47. Ishikawa HI, Nakane MT, Yoshizo EIM. Method
for thermochemical production of hydrogen from
water. U.S. Patent 3996342, 7 December 1976.
48. NASA Technical Reports Server (NTRS). Production
of Hydrogen. Id: 19740003618. Available from:
http://ntrs.nasa.gov/archive/nasa/casi.ntrs.nasa.gov/
19740003618_1974003618.pdf [accessed 26 December
2008].
49. Motupally S, Mah DT, Freire FJ, Weidner JW.
Recycling chlorine from hydrogen chloride. The
Electrochemical Society Interface Fall, 1998.
50. Aerojet-General Nucleonics San Ramon Ca,
Cesium-Water hydrogen production process for
the ammonia production feasibility study, Final
Technical Report, 2 October 196429 January 1965.
Available from: http://oai.dtic.mil/oai/oai?verb 5
getRecord&metadataPrex 5html&identier 5
AD0460841 [accessed 07 August 2009].
51. National Institute of Standards and Technology
(NIST). Available from: http://webbook.nist.gov/
chemistry/form-ser.html [accessed 26 February
2009].
52. The Exergoecology Portal. Available from:
http://www.exergoecology.com/excalc/ [accessed 26
February 2009].
53. McBride BJ, Zehe MJ, Gordon S. NASA Glenn
Coefcients for Calculating Thermodynamic
Properties of Individual Species, National Aeronautics
and Space Administration Glenn Research Center,
NASA/TP2002-211556, September 2002.
54. Cornelissen RL. Thermodynamics and sustainable
development. Ph.D. Thesis, University of Twente,
The Netherlands, 1997.
55. Rosen MA, Dincer I, Kanoglu M. Role of exergy in
increasing efciency and sustainability and reducing
environmental impact. Energy Policy 2008; 36:
128137.
56. Van Gool. Energy policy: fairly tales and factua-
lities. In Innovation and TechnologyStrategies and
Policies, Soares ODD, Martins da Cruz A, Costa
Pereira G, Soares IMRT, Reis AJPS (eds). Kluwer
Academic Publishers: Dordrecht, 1997; 93105.
57. Rosen MA, Scott DS. Comparative efciency
assessment for a range of hydrogen production
processes. International Journal of Hydrogen Energy
1998; 23(8):653659.
58. Yildiz B, Kazimi M. Efciency of hydrogen
production systems using alternative nuclear energy
technologies. International Journal of Hydrogen
Energy 2006; 31:7792.
Potential thermochemical and hybrid cycles I. Dincer and M. T. Balta
Int. J. Energy Res. 2011; 35:123137 r 2010 John Wiley & Sons, Ltd.
DOI: 10.1002/er
137

Vous aimerez peut-être aussi