You are on page 1of 16

Progress in Nuclear Energy 123 (2020) 103317

Contents lists available at ScienceDirect

Progress in Nuclear Energy


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

Review

Comparative review of hydrogen production technologies for nuclear hybrid


energy systems✩
Roxanne Pinsky a,b ,∗, Piyush Sabharwall b , Jeremy Hartvigsen b , James O’Brien b
a
University of Michigan, 2355 Bonisteel Blvd, Ann Arbor, MI, USA
b
Idaho National Laboratory (INL), Idaho Falls, ID, USA

ARTICLE INFO ABSTRACT

Keywords: Nuclear hybrid energy systems (NHES) have potential to capitalize on (1) producing multiple commodities,
Nuclear hybrid energy systems (NHES) i.e. electricity and hydrogen as well as (2) allowing for electricity grid load following, with hydrogen
Hydrogen production production during low electricity prices. Using nuclear thermal energy and electricity (from the reactor itself)
Electrolysis
makes hydrogen production an economically attractive option. The reactor can continuously operate at full
Thermochemical water splitting
capacity, sending excess heat and electricity towards hydrogen production, which could either be sold or
Hybrid water splitting
Steam methane reforming
converted back to electricity using fuel cells at high price times. Several hydrogen production technologies
exist, but in this study the focus is on processes that require heat and electricity. These candidates include
alkaline water electrolysis, proton exchange membrane (PEM) electrolysis, solid oxide electrolysis cells (SOEC),
thermochemical sulfur–iodine (S–I), calcium-bromide (Ca-Br) cycles, hybrid sulfur (HyS) and copper–chlorine
(Cu–Cl) cycles. Each have different minimum temperature requirements which can be coupled to Generation
III and IV reactor outlet temperatures: low (<300 ◦ C), medium (<750 ◦ C), and high (<950 ◦ C). Energy input
and material process flow diagrams were created for all technologies at compatible reactor temperatures
and compared to the most common commercially operating hydrogen production method: steam methane
reforming (SMR). Technology readiness levels (TRLs) and costs were also compared. The TRL of most systems
is still below commercial development, and hydrogen productions costs are still too high to be economic
without additional policy and/or other developments.

1. Introduction hydrogen is not readily available in nature, despite the fact that it is by
far the most abundant element in the universe, and must be produced
Hybrid nuclear energy systems utilize multiple energy resources to from a variety of different methods (Nikolaidis and Poullikkas, 2017).
couple multiple subsystems and produce energy commodities as out- These methods can include fossil fuel based hydrogen production,
puts (Antkowiak et al., 2012). This allows a nuclear energy system to biomass based hydrogen, and hydrogen from water, requiring heat and
operate at full capacity and either produce electricity for grid base-load electricity (Holladay et al., 2009).
power or use excess heat and electricity to produce a new commodity. Hydrogen may be safely stored and transported through conven-
As climate change advances due to increases in greenhouse gases, the tional means. It may be stored as a compressed gas, cryogenic liquid
need for renewable, cleaner alternatives to fossil fuels becomes more or solid hydride (Nikolaidis and Poullikkas, 2017). It can then produce
urgent. Pollutants from fossil fuels are already estimated to be high energy from fuel cells or used as a fuel for transportation, heating,
enough to affect public health and the environment specifically at high metal production (such as the reduction of iron oxide to iron), or pro-
density populations within the US (Holladay et al., 2009). Hydrogen
duction of gasoline, methanol, ethanol, ammonia, and other high value
(H2 ) fuel is a potential commodity that is clean, renewable and can be
chemicals (Holladay et al., 2009; Nikolaidis and Poullikkas, 2017). As
produced using thermal energy from a nuclear plant.
of 2017 the annual production of H2 reached 0.1 GT, used mainly for
Hydrogen is a promising alternative fuel (Holladay et al., 2009).
petroleum refining and treating metals, with a small fraction for fueling
It has the highest energy content compared to any other known fuel,
cars (Nikolaidis and Poullikkas, 2017). The demand for hydrogen fuel
as shown in Table 1 (Nikolaidis and Poullikkas, 2017). However,

✩ This work was supported in part by the U.S. Department of Energy, Office of Science, United States, Office of Workforce Development for Teachers and
Scientists (WDTS), United States under the Science Undergraduate Laboratory Internships Program (SULI).
∗ Corresponding author at: University of Michigan, 2355 Bonisteel Blvd, Ann Arbor, MI, USA.
E-mail address: rzpinsky@umich.edu (R. Pinsky).

https://doi.org/10.1016/j.pnucene.2020.103317
Received 16 November 2019; Received in revised form 11 February 2020; Accepted 24 February 2020
Available online 4 March 2020
0149-1970/Published by Elsevier Ltd.
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

Table 1 into hydrogen. This hydrogen may then be converted back to energy
Higher and lower (LHV) heating value of various fuels
during times of high electricity demand using fuel cells or be sold as
(Nikolaidis and Poullikkas, 2017).
its own commodity. Fig. 2 shows the routes a nuclear reactor may
Fuel HHV (MJ/kg) LHV (MJ/kg)
take for hydrogen production, depending on available reactor outlet
Hydrogen 141.9 119.9
temperature.
Methane 55.5 50
Ethane 51.9 47.8
As stated previously, there are a multitude of ways hydrogen may be
Gasoline 47.5 44.5 produced. The focus of this paper is on renewable hydrogen production
Diesel 44.8 42.5 from water that requires only excess heat and electricity that can
Methanol 20 18.1 be produced from the thermal energy of a nuclear reactor. These
water-based hydrogen production technologies are also compared to
the most common method of hydrogen production today — steam
methane reforming, which is used in approximately 48% of current
production (Department of Energy (DOE) Office of Energy Efficiency
& Renewable Energy, 2019). This method requires fossil fuels for
production and therefore releases greenhouse gases including carbon
dioxide (CO2 ).
The temperature of the water is also an important consideration.
Nuclear reactors have outlet temperatures that vary from approxi-
mately 300 to 950 ◦ C, depending on the design (Forsberg, 2019;
Goldberg and Rosner, 2011). Generation III reactors primarily consist
of Light Water Reactors (LWRs), which are the predominant reactor in
the United States. These are considered low temperature reactors and
have outlet temperatures estimated at 300 ◦ C. Generation IV reactors
are pushing on the average outlet temperature. For the purpose of
analysis, the LWR and Generation IV reactors have been split into
three categories: low (<300 ◦ C), medium (<750 ◦ C), and high (<950
Fig. 1. Influence of renewable energy shown in California electricity prices over a ◦ C, though realistically likely <900 ◦ C) outlet temperatures. The outlet
day (Forsberg, 2019).
temperature can determine what hydrogen production methods are fea-
sible. Table 2 includes a summary of the three temperature categories,
including reactor type, outlet temperature, and average plant size. Note
is expected to increase over the next few decades (Nikolaidis and that the conversion from thermal to electric power is estimated at 40%
Poullikkas, 2017; El-Shafie et al., 2019). for this and all further analysis. This efficiency could be achieved by
Coupling hydrogen production to a nuclear energy system offers advanced reactors, whereas for the LWRs the efficiency varies between
a multitude of advantages. Politically, the inconsistency of importing 30 to 33%. Also note that this is not an exhaustive list of all the
fossil fuels can lead to increases in prices and uncertainty in energy sup- potential concepts, but shows that reactors are being designed in all
ply. Environmentally, hydrogen does not emit greenhouse gases when three different temperature ranges.
combusted or produced from greenhouse gas free sources (i.e. produced This paper offers an overview of hydrogen production technologies,
using heat and electricity from solar or nuclear and producing hydrogen with a focus on water-based hydrogen production methods. It also
from water or biomass). Hydrogen has the potential to provide clean, includes an overview of storage and potential uses of hydrogen, such
domestic energy for increased economic and energy security (Holladay as for fuel or conversion back to electricity. Finally, a comparison
et al., 2009). is made on several hydrogen production methods, including steam
Current nuclear reactors operate for base load electricity produc- methane reforming, thermochemical water splitting, water electrolysis,
tion, which accounts for 60% of electricity demand (Antkowiak et al., and hybrid technologies. This comparison includes the technological
2012). However, the other 40% requires systems that can operate readiness levels, hydrogen costs, and heat and electricity requirements
for load-following conditions. This is because renewable energies such for all methods. It also includes a comparison of methods for high
as solar and wind peak at certain times (i.e. solar increases substan- (<950 ◦ C), medium (<750 ◦ C), and low (<300 ◦ C) water temperature
tially during the middle of the day), which can lead to volatile daily inputs to indicate possible hydrogen production methods for several
prices (Forsberg, 2019). The prices can fluctuate to negative values reactor classifications.
during high renewable capacities (Forsberg, 2019). Fig. 1 shows the
negative electricity prices arising from increased renewable (solar) 2. Hydrogen production technologies overview
energy in California over the course of a day (Forsberg, 2019). In these
cases, capacity payments may be required to keep generator systems A brief review of current and future hydrogen production methods
online when there is low wind or solar input (Forsberg, 2019). As re- is discussed. The majority of current hydrogen production technology
newable energy increases, the fraction of base-load energy will shrink, comes from fossil fuel production. 96% of commercial hydrogen comes
and more effort will be placed load-following technologies. Nuclear from a combination of natural gas (48%), heavy oils and naptha (30%),
systems alone are not economically effective when participating in this and coal (18%), which are mainly reformed using steam methane
load following structure, because they would not consistently able to reforming (SMR) (Nikolaidis and Poullikkas, 2017; Department of En-
operate at full capacity. Thus, producing other commodities, such as ergy (DOE) Office of Energy Efficiency & Renewable Energy, 2019).
hydrogen, will provide another lucrative option and make the whole The other 4% of hydrogen production comes from water electrolysis
system more economically attractive. using alkaline electrolyzers, a renewable technology (Zeng and Zhang,
The nuclear reactor coupled to hydrogen production has the poten- 2010). Hydrogen production can be divided into production from fossil
tial for higher overall energy use efficiency and better utilization of fuels and production from renewable resources. Renewable resources
capital equipment (Antkowiak et al., 2012). This is because the nuclear include both biomass hydrogen production and water splitting. Table 3
reactor can operate at full capacity, but also follow the electricity grid includes a summary of current and developed hydrogen production
as it varies throughout the day. When electricity prices are low, or technologies, feedstock, operating temperature range, efficiency, and
negative, the reactor can use a subsystem to convert heat or electricity developmental stage.

2
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

Fig. 2. Possibilities for nuclear hybrid hydrogen generation, including electricity for grid power and hydrogen production.

Table 2
Categorization of low, medium, and high temperature reactors for analysis (Goldberg and Rosner, 2011; Boardman, 2017; El-Emam and Özcan, 2019;
Yildiz and Kazimi, 2006).
Outlet temp (◦ C) Power𝐴𝑣𝑔 Power𝐴𝑣𝑔 Reactor Temp (◦ C) Power
(MWt) (MWe) (MWt)
Low (<300 ◦ C) 2500 1000 Light Water Reactor (LWR) 300 2500
Advanced Gas Reactor (AGR) 650–750 1550
Sodium Cooled Fast Reactor (SFR) 450–550 1250a
Medium (<750 ◦ C) 1800 720 Lead Cooled Fast Reactor (LFR) 480–570 1350a
Super Critical Water Reactor (SCWR) 500–625 2500
Molten Salt Fast Reactor (MSFR) 650–750 2500
Very High Temperature Reactor (VHTR) 900–950 600
High (<950 ◦ C) 1400 560 Gas Cooled Fast Reactor (GFR) 750–850 1450a
Molten Salt Reactor (MSR)b 750–950 2000
a Averaged from multiple reactor designs.
b
Advanced High Temperature Reactor (AHTR) design.

Table 3
Summary of current and future hydrogen production technologies. Note that ‘‘near term’’ estimates less than 5 years, ‘‘medium term’’ 5–10 years, and ‘‘long term’’ greater
than 10 years to commercial maturity (Holladay et al., 2009; Nikolaidis and Poullikkas, 2017; El-Shafie et al., 2019; Rashid et al., 2015; Dagle et al., 2017; Schmidt
et al., 2017).
Technology Production sub-method Feedstock Temperature range (◦ C) Efficiency (%) Maturity
SMR Hydrocarbons 700–1000 74–85 Commercial
POX Hydrocarbons 800–1000 60–75 Commercial
ATR Hydrocarbons 700–1000 60–75 Near term
Fossil fuel based Pyrolysis Hydrocarbons 1000–1400 51 Near term
Plasma reforming Hydrocarbons 900–1300 9–85 Long term
Aqueous phase reforming Carbohydrates 220–270 35–55 Medium term
Ammonia reforming Ammonia 800–900 28.3 Near term
Biomass gasification Biomass 800–1000 35–50 Commercial
Photolysis Water + Sunlight Ambient 0.5 Long term
Dark fermentation Biomass Ambient 60–80 Long term
Photo fermentation Biomass + Sunlight Ambient 1.9 Long term
MEC Biomass + Electricity Ambient 78 Long term
Renewable based Alkaline electrolysis Water + Electricity 40–90 62–82a Commercial
PEM electrolysis Water + Electricity 20–100 67–82a Commercial
SOEC Water + Electricity + Heat 700–1000 <110a Medium term
Thermochemical
Water + Heat 500–1000+ 20–45 Long term
Water splitting
PEC/Photoelectrolysis Water + Sunlight Ambient 12.4 Long term
a
Voltage efficiency (%) based on HHV of hydrogen (may be greater than 100%).

2.1. Production from fossil fuel reforming carbon monoxide (CO), CO2 , and other byproducts. Methane fuel pro-
cessing (from natural gas) is the most common commercial hydrogen
Fossil fuel hydrogen production technologies are commercially production method (Holladay et al., 2009). The biggest challenge to
available or developed and cheap due to low fossil fuel costs (Nikolaidis
hydrocarbon reforming processes comes from the catalyst poisoning
and Poullikkas, 2017). The main fossil fuel technology is hydrocarbon
reforming, in which a hydrocarbon fuel, such as natural gas and coal, due to sulfur, which is contained in most hydrocarbon fuels (Holladay
is converted through various reforming technologies into hydrogen, et al., 2009).

3
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

The main hydrocarbon reforming processes consist of SMR, partial can be combined with the reforming reactor for simplicity (Holladay
oxidation (POX), and autothermal reforming (ATR). In the reform- et al., 2009). Future research efforts are focused on improving catalyst
ing process, hydrocarbons react with either steam in an endothermic technology to prevent methanation and decreasing reactor size (due to
process (in SMR process), oxygen in an exothermic reactor (in POX pro- moderate space time yields) (Holladay et al., 2009).
cess), and both reactions combined for ATR (Nikolaidis and Poullikkas,
Ammonia reforming utilizes ammonia instead of hydrocarbons (Hol-
2017). The reactions all produce CO, so they all undergo a water-gas-
laday et al., 2009). Ammonia is made from nitrogen and hydrogen
shift reaction to convert the CO into CO2 and H2 (see SMR section
(NH3), and the reforming allows for the recovery of previously stored
for details). SMR has the advantages of extensive commercialization,
hydrogen. It has been proposed as a method for portable fuel cell
no pure oxygen required (only steam), lower process temperatures,
and high H2 to CO ratios relative to the other hydrocarbon reforming applications due to the extensive pipelines of ammonia that already
methods (Holladay et al., 2009). However, SMR has the highest air exist (as ammonia is a common fertilizer) (Holladay et al., 2009).
emissions. A further description of the SMR process is provided in the Ammonia reforming has low cost and a low environmental impact (El-
following section. POX, also a commercially available technology, has Shafie et al., 2019). It may be directly coupled with a high temperature
been proposed to produce hydrogen for automobile fuel cells and com- fuel cell, or decomposed in a plasma reactor (El-Shafie et al., 2019).
mercial applications (Nikolaidis and Poullikkas, 2017). This process has
a slightly lower efficiency than SMR based on methane’s higher heating
value (see Table 3). Since POX consists of a reforming reaction that is 2.1.1. Steam Methane Reforming (SMR)
exothermic, coke and hot spot formation are prevalent which makes SMR is the largest, most commercially developed hydrogen produc-
thermal management difficult (Holladay et al., 2009; El-Shafie et al., tion technology. For this reason, an in-depth process is described and
2019). Additionally, POX is capital intensive since the process requires this method will be compared to potential nuclear hybrid technologies.
pure oxygen (Nikolaidis and Poullikkas, 2017). However, POX is the An overview of this process is shown in Fig. 3.
best hydrocarbon reforming process to produce hydrogen from heavy Natural gas or other methane containing hydrocarbons, such as
feedstocks such as coal and heavy oil residues (Nikolaidis and Poul-
ethane, propane, butane, pentane, or light and heavy naptha are first
likkas, 2017). Autothermal reforming combines the benefits of both
sent to either a steam system to combust, heating steam, or a sulfur
SMR and POX: it uses heat from the exothermic POX reaction to assist in
removing device (Nikolaidis and Poullikkas, 2017). The sulfur remov-
the endothermic SMR reaction to increase the H2 /CO ratio (Nikolaidis
and Poullikkas, 2017). ATR is not at the commercial stage yet but ing process is important because sulfur can poison catalysts in the
will have lower investments costs than SMR or POX, a lower operating reformer and be detrimental to the reactions. Sulfur removal is typically
temperature than POX, and the potential for rapid stop and start times done either with chemical reactions or adsorptive technology. The main
when it is (Holladay et al., 2009; Nikolaidis and Poullikkas, 2017). commercial approach is the use of hydrodesulfurization (HDS) chemical
Other fossil fuel-based hydrogen technologies include pyrolysis, reactions (Holladay et al., 2009). In HDS, catalysts hydrogenate the
plasma reforming, aqueous phase reforming, and ammonia reforming. sulfur bearing molecules to release H2 S (Holladay et al., 2009). A
None of these technologies have reached commercial maturity. Pyroly- common adsorptive technique is using activated carbon to absorb the
sis is the process in which a hydrocarbon is decomposed into hydrogen sulfur (Holladay et al., 2009).
and carbon in the absence of water or oxygen. It can be done with Steam from the steam system and clean methane (CH4 ) then go
biomass/organic material and is used for producing carbon nanotubes to a reformer, where a catalyst activates the following endothermic
and spheres (Holladay et al., 2009). Pyrolysis has the advantage of
reaction:
simplicity, as it does not require processing steps such as water gas shift
and CO2 removal, making it cheaper to produce hydrogen (Nikolaidis CH4 + H2 O → CO + 3H2 (1)
and Poullikkas, 2017). However, any water or oxygen in the process
may be detrimental causing excess CO and CO2 emissions, and the Catalysts may be non-precious metals (such as nickel) or Group
H2 separation process may be expensive due to the low proportion VIII precious metals such as platinum or rhodium (Holladay et al.,
of hydrogen in the mixture (Holladay et al., 2009; Nikolaidis and 2009). Nickel catalysts are typically used in industry because they
Poullikkas, 2017). do not limit the reaction and are cheaper (Holladay et al., 2009).
Plasma reforming includes the same reactions as hydrocarbon re- Other alkaline components may be added to the catalyst to suppress
forming, but energy and radicals are provided by a plasma (Holladay
coke formation (Holladay et al., 2009). Additionally, to suppress coke
et al., 2009). Developed plasma reforming technology has the potential
formation the reaction takes place at high temperatures (700–1000 ◦ C),
to decrease costs, decrease deterioration of catalysts, decrease reactor
pressures up to 3.5 MPa and steam to carbon ratios of 3.5 (Nikolaidis
size and weight, increase response time, and increase the ability to
and Poullikkas, 2017). Once reformed, the syngas made of CO and H2
effectively reform heavy hydrocarbons (Holladay et al., 2009). Since
only free radicals are increased in temperature, overall reaction tem- are conditioned in a water-gas-shift reactor. Here, water and CO are
peratures may be lower. However, high electricity requirements and reacted with a catalyst to create more hydrogen and CO2 (Department
electrode erosion prove challenging for this technology (Holladay et al., of Energy (DOE) Office of Energy Efficiency & Renewable Energy,
2009). Plasma reforming is subdivided into two categories: thermal 2019). This is shown in the following reaction:
and non-thermal plasma. Thermal plasma involves raising temperature
CO + H2 O → CO2 + H2 (2)
of both free radicals and neutral species to elevated temperatures,
and therefore has high power requirements (El-Shafie et al., 2019).
The CO2 and other impurities are removed in a process called
Non-thermal plasma reforming involves increasing only the temper-
pressure-swing adsorption to leave nearly pure hydrogen (Department
ature of electrons, keeping neutral gas temperature low. There are
of Energy (DOE) Office of Energy Efficiency & Renewable Energy,
a few non-thermal plasma techniques: gliding arc plasma, dielectric
barrier discharge, microwave plasma, and corona discharge, and these 2019). Excess water is treated and cooled, and may be recycled back
techniques may see efficiencies anywhere between 9 and 85%. into the steam system. Additionally, CO2 may be captured via a carbon
In developed aqueous phase reforming, hydrogen is produced from capture and storage system, where the CO2 may be stored in a geologic
oxygenated hydrocarbons or carbohydrate at low temperatures (220– reservoir or under water to reduce greenhouse gas emission to the at-
270 ◦ C), and pressures of 25–30 MPa (Holladay et al., 2009). Because mosphere. Based on the HHV of methane, SMR has efficiencies between
this technique operates at low temperatures, a water gas shift reactor 74% and 85% (Nikolaidis and Poullikkas, 2017).

4
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

Fig. 3. Steam Methane Reforming (SMR) process (McKellar, 2010).

2.2. Production from renewable resources for a continuous stream of hydrogen to be produced (Nikolaidis and
Poullikkas, 2017). Alternatively, photo fermentation (also known as
Although hydrogen production from renewable resources is cur- photosynthetic bacterial hydrogen production), utilizes purple non-
rently low, it is expected to dominate over conventional technologies sulfur bacteria on the nitrogenase functionality (Holladay et al., 2009).
in the long term (Nikolaidis and Poullikkas, 2017). The two main Similar to photolysis, the bacteria produce hydrogen with sunlight,
methods of producing hydrogen come from biomass and water splitting. though in this case through fermentation. The advantage of photo fer-
Biomass is the second largest renewable resource, under hydropower, mentation over photolysis is that oxygen does not inhibit the reaction,
in the United States and will likely become the main substitute to however, the fermentation process is slow and requires a relatively high
petroleum (Holladay et al., 2009). Biomass consists of a wide array amount of energy (Holladay et al., 2009).
of sources such as animal residue, plant residue, sawdust, aquatic Microbial electrolysis cells (MEC) decompose biological material
plants, waste paper, grasses, wood, and much more (Holladay et al., directly through electrohydrogenesis. MEC is a modified microbial fuel
2009; Nikolaidis and Poullikkas, 2017). CO2 is released during biomass cell, where an external voltage is applied to the cell, in an anaerobic
hydrogen production, but only in the amount absorbed by organisms state, for acetate substrate decomposition (Holladay et al., 2009). The
while they were still living, which is considerably smaller in magnitude minimum theoretical applied voltage for this process is 0.11 V, but
than fossil fuels (Nikolaidis and Poullikkas, 2017). practically is likely greater than 0.3 V due to ohmic resistance and
Biomass hydrogen production processes can be further subdivided electrode overpotentials (Holladay et al., 2009). Complexity is high in
into biological and thermochemical processes. Thermochemical pro- these cells due to excess methane produced that must be intermittently
cesses include biomass gasification and others, such as pyrolysis, com- removed (Holladay et al., 2009).
bustion, and liquefaction. Biological processes include photolysis, dark Finally, water splitting processes can be divided into three group:
and photo fermentation, and microbial electrolysis cells. Thermochem- water electrolysis, thermochemical water splitting, and photoelectro-
ical biomass production is predominantly gasification, a mature and chemical water splitting (PEC), also known as photoelectrolysis. Water
commercially available technology (Holladay et al., 2009). In gasifi- splitting is the cleanest hydrogen production method as there are no
cation, biomass is converted into a hydrogen-rich gas in the presence greenhouse gases emitted. These processes all utilize some method to
of oxygen or air with the use of catalysts (Nikolaidis and Poullikkas, split water in the following reaction:
2017). Gasification produces methane and CO, which can be further
2H2 O → 2H2 + O2 (3)
converted into hydrogen with SMR and water gas shift reactors (Niko-
laidis and Poullikkas, 2017). Biomass gasification has relatively low Water electrolysis is the process where a potential difference is
thermal efficiencies due to moisture that must be vaporized in the placed between two electrodes, in an electrolyte, to break apart water
biomass (Holladay et al., 2009). It also can suffer from increased into its components. Electrolysis is the reverse process of the fuel cell,
costs due to the necessity of cleaning large amounts of tar produced in which hydrogen and oxygen are combined to produce electricity and
in gasification as well as large amounts of resources spent gathering water. Electrolysis requires a source of water, which is inexpensive,
biomass for commercial-scale production (Holladay et al., 2009). and electricity. Commercial products exist, which are made of alkaline
Biological processes, on the other hand, are still early in the de- electrolyzers (Holladay et al., 2009). Alkaline electrolysis breaks liquid
velopmental stage. They use organic material (micro-organisms) that water apart. Developed electrolysis methods include proton exchange
consume water and produce hydrogen. Biological processes have the membrane (PEM) electrolysis, which also operates in the liquid water
main disadvantage of low production and yield rates (Nikolaidis and region, and solid oxide electrolysis cells (SOEC), or high temperature
Poullikkas, 2017). Researchers continue to discover and develop micro- electrolysis, which operates with steam (700–1000 ◦ C) (Nikolaidis
organisms to advance biological processes (Holladay et al., 2009). and Poullikkas, 2017). High temperature electrolysis is less expensive
Within biological hydrogen production, there is photolysis, dark fer- than other electrolysis methods. In SOEC, part of the electrical energy
mentation, photo fermentation, and microbial electrolysis cells. Photol- requirement is replaced with thermal to achieve the required steam
ysis utilizes the fact that some micro-organisms, such as green algae and temperature. High electrolysis costs, due to the electricity requirement,
cyanobacteria, produce hydrogen with excess solar energy (Holladay currently prevent electrolysis from dominating hydrogen production
et al., 2009). It only requires water and sunlight, which is inexpensive on a large scale (Nikolaidis and Poullikkas, 2017). High pressure elec-
and globally available (Holladay et al., 2009). However, a large sur- trolyzers are currently being developed to reduce costs of compressing
face area is required to collect enough sunlight, and organisms cease hydrogen post-electrolysis (Holladay et al., 2009). Due to the poten-
producing hydrogen in the presence of oxygen (a by-product of water tial of electrolysis methods in production of H2 , an in-depth process
splitting) (Holladay et al., 2009). Efficiencies are also theoretically diagram of each electrolysis method is studied in this effort.
limited to less than 1% (Holladay et al., 2009). Thermochemical water splitting involves using only heat to split
Dark fermentation uses anaerobic bacteria, in the dark, under oxy- water. Based on the Gibbs free energy to split water in a single step
gen free conditions, to ferment carbohydrate rich substrates (Nikolaidis Eq. (3), the minimum temperature required to create hydrogen is over
and Poullikkas, 2017). Pure, simple sugars such as glucose is a pre- 1700 ◦ C for partial dissociation, and over 4000 ◦ C for total dissoci-
ferred substrate, but is relatively expensive and unavailable in large ation (Grimes et al., 2007). This temperature is extremely high for
quantities (Nikolaidis and Poullikkas, 2017). An advantage of dark most heat sources, so there has been heavy development in creating
fermentation is that space is not an issue, as no light is required multiple step cycles to reduce reaction temperatures. These cycles

5
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

produce hydrogen and oxygen and recycle the chemicals through the stage, like the SOECs. An additional problem is poisoning via CO2 (due
steps. Thus far, over 100 cycles in over 800 literature papers have been to impure inputs), which can include a cost penalty. AFC and PEMFC
identified, operating at a multitude of temperature ranges at various are susceptible to this CO2 poisoning (Gielen and Simbolotti, 2005).
stages of development (Rosen, 2010). The lower temperature cycles SOFC has fewer problems with impurities, but faces a wide array of
may achieve minimum temperatures around 500 ◦ C while others have materials challenges due to their high operating temperatures (700–
a maximum temperature over 1000 ◦ C. Cycles of interest to nuclear 1000 ◦ C) (Rand and Dell, 2007). The overall efficiency of converting
hybrid technology will be further discussed in the Nuclear Hybrid energy from a nuclear plant to hydrogen and back to electricity through
Candidates section. These cycles are more developed and have been a fuel cell could be as low as 12% due to energy conversion losses, elec-
hypothesized to work with reactors already, including the sulfur–iodine trolyzer and fuel cell efficiencies, and losses in hydrogen compression
and calcium–bromine cycles (Yildiz and Kazimi, 2006; McKellar, 2010). and electricity transmission/distribution (Rand and Dell, 2007).
Additionally, beyond a simple thermochemical cycle involving only Fuel cells may be supplied by pipelined hydrogen or directly from
water and heat, hybrid processes where one step involves electrolysis hydrogen on site (Rand and Dell, 2007). The hydrogen generated in a
have also been developed. The main hybrid processes in focus are the nuclear plant may have a fuel cell directly attached to produce elec-
hybrid sulfur cycle and copper–chlorine cycle. tricity to the grid during times of high electricity costs. Alternatively,
PEC or photoelectrolysis is another developed process which only the hydrogen produced at a nuclear plant may be pipelined and sent to
requires water and sunlight. In PEC, sunlight is absorbed in semicon- fuel cells in other areas to produce heat and/or electricity at a specific
ductor materials, and if the photon has an energy greater than the location.
bandgap, and electron–hole pair is created which splits water into its Storage of hydrogen is generally simpler for energy storage ap-
components (Nikolaidis and Poullikkas, 2017). The minimum voltage plications when compared to energy carrying (propulsion) applica-
required is 1.23 V, based on the Gibbs free energy of liquid water. A tions (Rand and Dell, 2007). It also is cheaper to store bulk hydrogen
high bandgap energy is necessary to split the water which can decrease from large centralized plants (up to four times cheaper than distributed
overall efficiency (Nikolaidis and Poullikkas, 2017). Efficiency is also plants, approximately 100 times smaller) (Rand and Dell, 2007).
decreased due to imperfections in semiconductor crystalline struc- To store hydrogen, it must be compressed, typically anywhere from
ture (Holladay et al., 2009). Additionally, high costs are involved given 20 to 80 MPa. This may be done adiabatically, isothermally, or in a
limitations in semiconductor materials and efficiencies (Nikolaidis and multi-stage process (Rand and Dell, 2007). A multi-stage process will
Poullikkas, 2017). likely require an energy input of around 10% of the HHV to compress
All hydrogen production technologies previously discussed are to 50 MPa, corresponding to potentially large ( 20%) electrical en-
shown in Table 3. Although any of these processes that require either ergy losses in the system (Rand and Dell, 2007; Bossel and Eliasson,
heat or electricity could be coupled to a nuclear reactor, the main 2003). Adiabatic compression requires more energy than multistage,
interest is generating clean, efficient energy when coupled to the
and isothermal requires less for all pressures (Bossel and Eliasson,
reactor. Therefore, the main interest is in water splitting, specifically
2003).
electrolysis and thermochemical water production.
There may be diffusion problems with compression, as H2 is the
smallest molecule and therefore has relatively high diffusivity. Al-
3. Hydrogen for energy storage
ternative storage methods include liquefaction, which can have high
energy penalties (21% of HHV for large capacity plants (greater than
Once hydrogen is produced, it may either be used for energy storage
1000 kg H2 /hr), and up to 110%, greater than the HHV of hydrogen,
and eventually converted back to heat or electricity, or as an energy
for advanced 1 kg/hr plants), storing hydrogen in a metal hydride
carrier, as shown in Fig. 2. For energy storage, fuel cells are the
(which has several pending heat transfer engineering problems), or
solution. Fuel cells are clean sources of flexible power, either heat or
chemical storage (which may be expensive or require high tempera-
electricity, are silent, and have relatively high energy efficiencies (Rand
tures) (Rand and Dell, 2007; Gielen and Simbolotti, 2005; Bossel and
and Dell, 2007; Gielen and Simbolotti, 2005). Fuel cells include an
Eliasson, 2003). However, economies of scale exist that can reduce
electrolyte to conduct ions, a positive and negative electrode, a plate
liquefaction capital costs by a factor of 4 by increasing production by a
to connect cells, and current collectors. Stacks of fuel cells are used
factor of around 20 (Gielen and Simbolotti, 2005). Fig. 4 compares the
to generate a desired amount of power, ranging from a few watts to
energy requirements for compression to 20 and 80 MPa adiabatically,
multi-MW, which allows them to be extremely versatile.
isothermally, and in a multi-stage process with liquefaction energy
Fuel cells are essentially the opposite of electrolysis, as it takes
penalties (Bossel and Eliasson, 2003).
in oxygen and hydrogen and produces water and power. In fact, de-
velopment of fuel cells is beneficial to both technologies. As shown
in the Nuclear Hybrid Candidates section on water electrolysis, fuel 4. Hydrogen as energy carrier
cells may simply run in reverse of electrolyzers. Alkaline electrolyte
(AFC), proton-exchange membrane (PEMFC) fuel cells, and solid oxide Hydrogen has been used as an energy carrier to propel spacecraft
fuel cells (SOFC), all share similar names, properties, and temperature since the 1960s. Work is being developed to use hydrogen for other
ranges as their electrolysis counterparts. Alternative fuel cells exist that propulsion needs, such as transportation. Hydrogen cars offer a clean
can take in non-hydrogen fuel, such as the direct methanol fuel cell alternative to vehicles powered by petrol or diesel but is not cur-
(DMFC) which takes methanol as input (Rand and Dell, 2007). rently competitive due to lower performances (reduced efficiencies)
For energy storage, the AFC, PEMFC, and SOEC are ideal for and higher prices (Rand and Dell, 2007).
nuclear-hydrogen systems. This is because they use the same materials Hydrogen as a fuel is low density (0.09 kg/m3 ) at standard temper-
as their electrolysis cells. Therefore, electrolysis can take place in the ature and pressure. It also has a low boiling point of 20.3 K (and liquid
same cell during times of low electricity, hydrogen and oxygen (or air) density of 70.8 kg/m3 ), requiring sophisticated technology to contain
stored for future use, and then be used as fuel cells to create electricity in this state. Hydrogen also has the highest standard heat of formation
to supply to the grid. AFC and PEMFC, both of which have been compared to other conventional fuels (higher heating value HHV) of
commercially developed (though less developed for PEMFCs) report 141.9 MJ/kg. However, due to the low density of hydrogen, higher
low start-up times (less than 10 min), enabling a quick switch to energy volumes are required to reach HHV densities compared to compressed
production (Rand and Dell, 2007). However, the SOFC still reports or liquified methane or petrol (Rand and Dell, 2007). For comparison,
long start up times (1-5 h), which is a potential problem for quick methane has a density of 0.72 kg/m3 and a HHV of 55.5 MJ/kg (Bossel
switches (Rand and Dell, 2007). SOFCs are still in the developmental and Eliasson, 2003).

6
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

low-carbon steel. This is an expensive task (Rand and Dell, 2007). Hy-
drogen pipelines do currently exist, with several thousand km currently
installed (Gielen and Simbolotti, 2005). Hydrogen fueling stations also
currently exist, with 115 in the US as of 2005 (Rand and Dell, 2007).
However, if hydrogen became widely used as a fuel for cars, new
infrastructure would need to be in place, which could be 40%–50%
more expensive than natural gas infrastructure (Rand and Dell, 2007).
Additionally, hydrogen pipelines require significantly more energy
per unit volume than natural gas pipelines, mainly for continual com-
pression of the fuel. Around 1.4% of hydrogen flow in a pipe is lost
per 150 km to energize compressors, compared to 0.3% of methane
(natural gas) for the same energy flow (Bossel and Eliasson, 2003).
This is an energy increase from natural gas pipelines by a factor of
4.7. The volumetric energy density of hydrogen is low, so for the same
energy flow, 3.1 times the volumetric flow, and therefore more energy,
is required compared to natural gas (Bossel and Eliasson, 2003). This
is calculated from the volumetric HHV of hydrogen and methane, 12.7
MJ/m3 and 40.0 MJ/m3 , respectively.
Fig. 4. Energy comparison (as a % of H2 HHV) for several hydrogen storage
methods (Bossel and Eliasson, 2003). 5. NHES hydrogen candidates

Reactor temperatures range from 300–950 ◦ C for Generation III and


Hydrogen has several drawbacks as a fuel source. It can be excep- IV reactors. These reactors may utilize excess heat and electricity dur-
tionally dangerous if leaked into an enclosed space, as it takes only ing off peak times for the production of hydrogen. The main interests
1/14th the energy needed to ignite compared to natural gas, and the in producing hydrogen fall under water electrolysis, thermochemical
flame is nearly invisible (Rand and Dell, 2007). Special, expensive water splitting, and a hybrid of the two systems. The next sections
cryostat units are needed for liquid fuel storage with vacuum and multi- will include background, processes, and assumptions used for modeling
level insulation (Rand and Dell, 2007). The cost of this specialized these candidates at various temperatures and power levels.
equipment is one reason hydrogen has not advanced as a widely
popular fuel beyond use in space exploration (Rand and Dell, 2007). 5.1. Water electrolysis
However, given the use of hydrogen in space exploration, it is clearly
possible to technologically manage handling and storage problems. Water electrolysis, as mentioned previously, involves the decom-
Fuel cells are not exclusively used for energy storage. They have position of water into H2 and O2 Eq. (3). The reaction has specific
been used to convert energy on spacecraft since the 1960s (Rand energy requirements based on the temperature and pressure of the
and Dell, 2007). There have been significant efforts to develop fuel water. The Gibbs free energy change of water formation, 𝛥𝐺, enthalpy
cells for propulsion, specifically focusing on automobiles (Gielen and change of water formation 𝛥𝐻, and entropy change 𝛥𝑆 (multiplied by
Simbolotti, 2005). Fuel cells have also been proposed for energy appli- temperature T) determine the minimum theoretical required voltage
cations ranging from submarines and aircraft to consumer electronics to decompose water as a function of temperature and pressure (Santos
such as laptops. However, further development is necessary in the et al., 2013):
renewables sector of hydrogen production, otherwise hydrogen used
as an energy carrier will be produced from fossil fuels mitigating the 𝛥𝐺 = 𝛥𝐻 − 𝑇 𝛥𝑆 (4)
purpose of using hydrogen as a clean energy source (Rand and Dell,
Note that the reactions taking place at the anode and cathode are
2007).
dependent on the electrolysis type (materials based), but the overall
PEMFCs are the proposed fuel cell of choice for transportation
reaction leads to the same overall voltage requirement. Also note
initiatives (Gielen and Simbolotti, 2005). Currently, PEMFC stack costs
that, for simplicity of the electrolysis model, pressures are assumed
exceed US $2000/kW, which must be lowered to around $50/kW (auto-
to be 1 atm. It is known that efficiencies decrease slightly (due to
motive costs, not stationary generation costs) to be competitive (Gielen
increased voltage requirements) for higher pressure systems but can
and Simbolotti, 2005). It has been estimated that a current reduction to
lead to power savings because the hydrogen is produced at a higher
$100/kW is feasible (Gielen and Simbolotti, 2005). Fuel cells also need
pressure (Marangio et al., 2009). However, this analysis does not take
to improve durability and reliability before they are ready for commer-
into account energy required to compress and store hydrogen, only to
cial transportation (Gielen and Simbolotti, 2005). The goal is mobile
produce it through various methods.
operation of 3000–5000 h for cars and 20,000 h for buses (Gielen and
Simbolotti, 2005). The Gibbs free energy determines the equilibrium cell voltage
Storage of hydrogen for energy carrying applications are important, 𝐸𝐸𝑞𝑢𝑖𝑙𝑖𝑏𝑟𝑖𝑢𝑚 , or the absolute minimum voltage required for electrolysis
because the storage units must be small and light enough to be on to occur, as shown in Eq. (5) (Zeng and Zhang, 2010):
the vehicle. Currently, only the compressed gas storage method is 𝛥𝐺 = 𝑛𝐹 𝐸𝐸𝑞𝑢𝑖𝑙𝑖𝑏𝑟𝑖𝑢𝑚 (5)
practical for hydrogen fueled automobiles. Developments in composite
containers, as opposed to steel, have been made but still have problems For water, n, the number of moles of electrons transferred in the
with high costs, energy penalties from compression, and safety issues reaction is 2 and F is the Faraday constant, at 96485 C/mol. At 25 ◦ C,
(rapid loss of hydrogen) (Rand and Dell, 2007). The U.S., as of 2015, 1 atm (standard temperature and pressure), 𝛥𝐺 = 237.1 kJ/mol and
targeted 10.8 MJ/kg as a benchmark for vehicle storage as of 2015, a 𝐸𝐸𝑞𝑢𝑖𝑙𝑖𝑏𝑟𝑖𝑢𝑚 = 1.23 V. The equilibrium voltage describes the minimum
difficult target to meet (Rand and Dell, 2007). electrical energy demand for electrolysis. Fig. 5 shows how temperature
Pipelining hydrogen also has challenges. While the hydrogen can effects Gibbs free energy, enthalpy, and entropy (O’Brien et al., 2010).
go through natural gas pipelines, there may be significant material Note that there is a drop in energy demand once water becomes
degradation due to the use of high-carbon steel that hydrogen reacts a vapor. Also note that, at the equilibrium voltage, the reaction is
with. Instead, new pipeline would likely need to be laid, consisting of endothermic and will not proceed without an additional heat source.

7
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

𝑅𝑇 𝑖
𝜂𝐾 = 𝑎𝑟𝑐𝑠𝑖𝑛ℎ( ) (10)
𝛼𝐾 𝐹 2𝑖0,𝐾
Note that different authors cite exchange current densities orders of
magnitude different for the same electrode material and temperature,
as seen in Carmo et al. (2013). Therefore, the chosen 𝑖0 for the different
water electrolysis models is only one estimate and is assumed to be
constant with temperature.
Hydrogen and oxygen production rates can also be calculated based
on the total cell current I. The total current may be calculated based on
the total cell area and current density. The production rates are given
in Eqs. (11) and (12) (Marangio et al., 2009). The oxygen production
rate in mol/s is half as much as hydrogen (see Eq. (3)) but is 8 times
larger when converted to kg/s, given that the oxygen molecule is 16
times larger than hydrogen (approximately 32 g/mol compared to 2
g/mol).
Fig. 5. Energy demand for water electrolysis at various temperatures and 1 mol 𝐼
atm (O’Brien et al., 2010). H2,𝑃 𝑟𝑜𝑑𝑅𝑎𝑡𝑒 [ ]= (11)
s 2𝐹
mol 𝐼 H2,𝑃 𝑟𝑜𝑑𝑅𝑎𝑡𝑒
O2,𝑃 𝑟𝑜𝑑𝑅𝑎𝑡𝑒 [ ]= = (12)
s 4𝐹 2
Another important voltage to note is the thermoneutral voltage, The total energy consumed per normal volume (at standard temper-
𝐸𝑇 ℎ𝑒𝑟𝑚𝑜𝑛𝑒𝑢𝑡𝑟𝑎𝑙 . The thermoneutral voltage is the voltage required for no ature and pressure) of hydrogen can also be calculated in the following
heat to be added (it is the transition point between endothermic and equation (Guenot et al., 2015):
exothermic reactions). It remains relatively constant with temperature,
kWh 𝑛𝐹
as it depends mostly on the heat of formation of water (285.8 kJ/mol 𝑊𝑒 [ ]= 𝐸 = 2.393 ∗ 𝐸𝑐𝑒𝑙𝑙 (13)
Nm3 𝑡𝑉𝑚 𝑐𝑒𝑙𝑙
for liquid and 241.8 kJ/mol for gas at atmospheric pressure). It is
related to the enthalpy change in Eq. (6) (Zeng and Zhang, 2010): where 𝑉𝑚 is the molar volume of an ideal gas at 1 atm and 25 ◦ C, equal
to 0.0224 Nm3 (N stands for normal volume of ideal gas), and 𝑡 is the
𝛥𝐻 = 𝑛𝐹 𝐸𝑇 ℎ𝑒𝑟𝑚𝑜𝑛𝑒𝑢𝑡𝑟𝑎𝑙 (6) number of seconds in an hour for conversion (equal to 3600), 𝐹 is the
The thermoneutral and equilibrium voltages are good for describing Faraday constant, and 𝑛 is 2 for the electrolysis process.
theoretical minimum energy requirements, but they are lower than One important property of electrolysis, like with fuel cells, is the
the expected required voltage, 𝐸𝑐𝑒𝑙𝑙 . Real electrolysis cells have prob- ability to stack and scale beyond one cell. A total number of cells can
lems such as activation overvoltage or overpotentials on the cells and be estimated to give a total power and heat requirement and total
resistances throughout that must be accounted for. These loss terms hydrogen and oxygen production rates. Additionally, it should be noted
may increase energy consumption by around 1.5–2.2 times more than that there is a heat requirement, estimated with 𝑇 𝛥𝑆, as well as an
theoretical (Rashid et al., 2015). The cell voltage required is described electricity requirement from the cell voltage. For alkaline and PEM
in Eq. (7) (Zeng and Zhang, 2010). electrolysis, the heat requirement already exists in the water, since
these processes occur with liquid water, so only electrical energy must
𝐸𝑐𝑒𝑙𝑙 = 𝐸𝐸𝑞𝑢𝑖𝑙𝑖𝑏𝑟𝑖𝑢𝑚 + 𝜂𝐴 + 𝜂𝐾 + 𝑖𝑅𝑐𝑒𝑙𝑙 (7) be provided. For SOEC, however, water must be converted to steam
where 𝜂𝐴 is the anode overvoltage, 𝜂𝐾 is the cathode overvoltage, and and elevated to 700–1000 ◦ C, so the heat requirement may not be
𝑖 is the current density. 𝑅𝑐𝑒𝑙𝑙 is the sum of all area specific resistances neglected. Since nuclear reactors provide heat and electricity, this is
in the cell: electric, anode, cathode, bubbles (resistance due to partial not an issue for coupling SOEC.
coverage of electrodes by bubbles), and membranes (Santos et al., The last important set of equations to model electrolysis are process
2013). It is estimated to be somewhere between 0.01–2 Ω cm2 for efficiencies (Zeng and Zhang, 2010). There are a few efficiencies com-
simplicity, but the resistance is dependent on materials (electrolysis monly used. The faradic efficiency is the same as the voltage efficiency,
type) and temperature (Phillips et al., 2017; Siracusano et al., 2018; and describes the ratio of voltage needed to split water based on Gibbs
Hauch, 2008). The term 𝑖𝑅𝑐𝑒𝑙𝑙 (V) consists of the ohmic losses in the free energy to voltage needed for electrolysis:
cell. Resistance can be minimized by minimizing the interelectrode 𝐸𝐸𝑞𝑢𝑖𝑙𝑖𝑏𝑟𝑖𝑢𝑚
gap and optimizing the current density based on electrode surface 𝐹 𝑎𝑟𝑎𝑑𝑖𝑐 𝐸𝑓 𝑓 𝑖𝑐𝑖𝑒𝑛𝑐𝑦 % = ∗ 100 (14)
𝐸𝐶𝑒𝑙𝑙
morphology and electrolyte concentrations (Phillips et al., 2017). There
may also be a diffusion overvoltage that exists in the case of mass Thermal efficiency is another measure of how well the electrolysis
transport limitations (Marangio et al., 2009). This typically occurs at system works, and it measures the ratio of thermoneutral voltage,
high current densities. However, in this model, it is assumed that there which is based on the enthalpy change of water decomposition, to the
is a continuous supply of water, and the current density stays low cell voltage.
enough that mass transport limitations are not reached. 𝐸𝑇 ℎ𝑒𝑟𝑚𝑜𝑛𝑒𝑢𝑡𝑟𝑎𝑙
To find anode and cathode activation overpotentials, which are 𝑇 ℎ𝑒𝑟𝑚𝑎𝑙 𝐸𝑓 𝑓 𝑖𝑐𝑖𝑒𝑛𝑐𝑦 % = ∗ 100 (15)
𝐸𝐶𝑒𝑙𝑙
functions of the current density i, the Butler–Volmer equation applies,
Finally, the efficiency can also be measured in terms of the higher
and is shown in Eq. (8) (Marangio et al., 2009):
heating value (HHV) of hydrogen. It is the ratio of power output (taken
𝛼1 𝐹 𝛼 𝐹 as the amount of power gotten from the amount of hydrogen produced
𝑖 = 𝑖0 [𝑒𝑥𝑝( 𝜂 ) − 𝑒𝑥𝑝(− 2 𝜂𝑎𝑐𝑡 )] (8)
𝑅𝑇 𝑎𝑐𝑡 𝑅𝑇 and the HHV value) to power input for electrolysis (Kabza, 2010). This
where 𝑖0 is the exchange current density (which depends on electrode is typically the quotable efficiency as it measures how efficient the
material), 𝛼1,2 are transfer coefficients, and 𝜂𝑎𝑐𝑡 is the activation over- system is at producing hydrogen. In these models the assumed conver-
voltage (of either the cathode or anode). If the transfer coefficients sion efficiency from heat to electricity is 40%. (McKellar, 2010). The
are assumed to be equal, then the following expressions are used to hydrogen production efficiency, as shown in Eq. (16), is typically lower
find the anode (A) and cathode (K) overpotentials, where 𝛼𝐴 = 2 and than the thermal or faradic efficiency due to electricity conversion
𝛼𝐾 = 0.5 (Marangio et al., 2009): efficiencies.
𝑅𝑇 𝑖 𝐻𝐻𝑉 ∗ H2,𝑃 𝑟𝑜𝑑𝑅𝑎𝑡𝑒
𝜂𝐴 = 𝑎𝑟𝑐𝑠𝑖𝑛ℎ( ) (9) H2,𝑃 𝑟𝑜𝑑. 𝐸𝑓 𝑓 𝑖𝑐𝑖𝑒𝑛𝑐𝑦 % = ∗ 100
𝛼𝐴 𝐹 2𝑖0,𝐴 𝐼𝑉

8
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

285.83 ∗ 1000 [J∕mol] ∗ H2,𝑃 𝑟𝑜𝑑𝑅𝑎𝑡𝑒 [mol∕s]


= ∗ 100
𝐸𝑙𝑒𝑐𝑡𝑟𝑖𝑐𝑖𝑡𝑦 [W]∕0.4 + 𝐻𝑒𝑎𝑡 [W]
(16)
The equations as described above provide the foundation for model-
ing alkaline, PEM, and SOEC electrolysis. Each electrolysis process has
specific processes, temperatures, and materials that change the inputs
of ion exchange current density, running current density, and resistance
into the model. It is assumed that the electrode area for all cells is
225 cm2 based on the model from Bragg-Sitton et al. (2014). The total
power input is based on reactor size, and the total number of cells
necessary to utilize this output is not necessary for comparison to other
water splitting processes.

5.1.1. Alkaline electrolysis


Alkaline electrolysis, as stated previously, is the most commercially
developed electrolysis method. It operates in the temperature range
of around 40–90 ◦ C but has been quoted to operate as low as 20
◦ C (Rashid et al., 2015; Sapountzi et al., 2017). In alkaline electrolysis,

water is split at the cathode and OH−travels to the anode (Holladay Fig. 6. Schematic of Alkaline electrolysis.
et al., 2009), as shown in Eqs. (17) and (18).
Anode:

4OH− → O2 + 2H2 O + 4𝑒− (17) 5.1.2. Proton Exchange Membrane (PEM) electrolysis
Proton exchange membrane (PEM) electrolysis, also known as
Cathode: polymer-electrolyte membrane electrolysis, uses a polymer electrolyte
as opposed to a liquid as seen in alkaline electrolysis. It operates at
2H2 O + 2𝑒− → H2 + 2OH− (18)
similar temperatures, 20–100 ◦ C, and converts liquid water (Rashid
Hydrogen is produced at the cathode and oxygen at the anode. The et al., 2015). Recent developments have been investigated to oper-
product gases are separated by a diaphragm, which is permeable to ate at temperatures around 200 ◦ C for fuel cell applications, which
the OH−carrier as well as water (Carmo et al., 2013). The electrodes translate to electrolysis applications (Sapountzi et al., 2017). PEM uses
are both submerged in a liquid electrolyte. The electrolyte is a liquid perfluorosulfonic acidic polymer solid electrolytes which also act as the
and can either be potassium hydroxide KOH or sodium hydroxide membrane (Rashid et al., 2015). PEM electrolysis may be operated at
NaOH, where KOH is the most common (Zeng and Zhang, 2010). While high pressures, up to 40 MPa, which will lower the energy requirements
there are many electrode materials that work (such as cobalt or iron), for compressing hydrogen for storage (Gielen and Simbolotti, 2005).
nickel is the most popular due to its high activity and availability These membranes allow for the proton, the charge carrier, to pass
combined with low cost (Zeng and Zhang, 2010). The exchange current through while preventing other gases from moving. The reactions that
density 𝑖0 of nickel has been reported to be around 1.1*10−4 A/cm2 occur in PEM are shown in Eqs. (19) and (20) (Holladay et al., 2009):
at the cathode (in a 50% KOH electrolyte at 80 ◦ C) and 4.2*10−6 Anode:
A/cm2 at the anode (Miles et al., 1976). These were the values used
2H2 O → O2 + 4H+ + 4𝑒− (19)
to calculate anode and cathode overpotentials. Pressure operation can
range from atmospheric to 12 MPa for advanced electrolyzers (Gielen Cathode:
and Simbolotti, 2005).
2H + 2𝑒− → H2
+
(20)
The operational current density has been reported anywhere from
0.1 to 0.4 A/cm2 (Holladay et al., 2009; Phillips et al., 2017; Carmo Electrodes are typical noble catalysts, typically platinum Pt for
et al., 2013). For modeling purposes, the current density was chosen the cathode and iridium Ir for the anode, which both have rela-
to be 0.3 A/cm2 , as this was within the range quoted in the majority tively high activity (Rashid et al., 2015). Other electrode materials
of the literature. The resistance 𝑅𝑐𝑒𝑙𝑙 is difficult to calculate because it include rhodium, rubidium, palladium, gold, and oxides of these cata-
depends on the particular cell itself, but for the purpose of the model is lysts (Rashid et al., 2015). The exchange current density at the cathode
estimated to be 0.2 Ω-cm (Santos et al., 2013). Not much more water is estimated to be 1*10−3 A/cm2 and 1*10−7 A/cm2 at the anode (Choi
than that used to produce hydrogen is needed as cooling is limited, et al., 2004). A schematic of the overall PEM process is shown in Fig. 7.
given the low operational temperatures. The total water consumption PEM electrolyzers have low ionic resistances which allow them
has been estimated to be around 11.5 times the amount of hydrogen to operate at high current densities, anywhere between 0.6 and 3
produced (Genovese et al., 2009). A/cm2 (Holladay et al., 2009; Siracusano et al., 2018; Carmo et al.,
A schematic of the alkaline electrolysis is shown in Fig. 6, including 2013). For modeling purposes, the current density was set at 2 A/cm2 .
common materials. High current density operation with relatively low ohmic resistances
Some advantages of alkaline electrolysis include stable, pure hydro- lead to high efficiencies and compact production rates. Literature val-
gen production (though slightly less pure than PEM) with non-noble ues of ohmic resistance were found as a function of temperature and
catalysts, and mature technology (Carmo et al., 2013; Sapountzi et al., estimated to be 0.2 Ω-cm at 30 ◦ C, 0.1 Ω-cm at 60 ◦ C, and 0.08 Ω-
2017). Suppliers of this electrolysis technology include De Nora SAP, cm at 90 ◦ C (Siracusano et al., 2018). Like alkaline water electrolysis,
Nel, Hydrogenics, Teledyne Energy Systems and General Electric (Sa- the total water consumption was estimated around 11.5 times the
pountzi et al., 2017). It is also highly scalable, because the cells amount of hydrogen produced, as little cooling is necessary during the
are stackable. However, some disadvantages are corrosion from the process (Genovese et al., 2009).
electrolyte, gas permeation through the diaphragm, and slow dynam- There are several advantages of PEM electrolysis. The solid elec-
ics (Sapountzi et al., 2017). It can also suffer from limited current trolyte allows for compact design and, combined with the high reactiv-
densities and low pressure operation (in large scale units), forcing large ity of the electrodes, leads to fast response times (<10 min) (Boardman,
volume designs (Rashid et al., 2015; Carmo et al., 2013). 2017; Sapountzi et al., 2017). Highly pure hydrogen is also produced in

9
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

Fig. 7. Proton Exchange Membrane (PEM) electrolysis process. Fig. 8. Solid Oxide Electrolysis Cell (SOEC) or High Temperature Electrolysis (HTE)
process.

the process (Sapountzi et al., 2017). PEM cells are also highly scalable.
On the downside, acidic electrodes cause degradation, limiting opera- SOEC is a promising technology with high application flexibility,
tion lifetime, and the polymer membranes have high costs (Sapountzi high scalability, pure hydrogen (though less pure than other electrolysis
et al., 2017). These are the main limitations to current widespread com- methods due to residual steam) production, and enhanced reaction
mercialization. However, up to MW-scale PEMs exist, by the company kinetics and thermodynamics at higher temperatures (reducing the
Proton OnSite (OnSite, 2019). Although PEM’s have been commercial- equilibrium cell voltage) with low capital costs (Boardman, 2017;
ized to the MW-scale, GW-scale PEMs are still in the research and Sapountzi et al., 2017). Some disadvantages come in the form of
development stage with a technology readiness level (TRL) of 4. degradation of electrodes and ceramic electrolyte at high temperature
operation. This means low durability and lack of stability which must
5.1.3. Solid Oxide Electrolysis Cells (SOEC) be improved before SOEC can become commercially available (Rashid
SOEC operates close to the output temperatures of nuclear reactors, et al., 2015; Sapountzi et al., 2017). Hydrogen is also the least pure
which gives this technology a unique advantage over its ambient in this method (relative to electrolysis methods, due to residual steam
in the system given the oxygen charge carrier). There is a developing
temperature counterparts. It operates using steam and is sometimes
technology which uses a solid ceramic membrane with steam (500–
called high temperature electrolysis (HTE), as opposed to liquid water,
1000 ◦ C), with an H+ charge carrier like PEM, that can produce high
at temperatures between 700 and 1000 ◦ C (Rashid et al., 2015). A solid
purity hydrogen (Sapountzi et al., 2017). This technology is still in
electrolyte is also used, making SOEC compact and have quick response
early development and will not be considered for further analysis,
times like PEM electrolysis, though in this case the electrolyte is a
but is a potential solution to produce pure hydrogen at lower steam
ceramic. It is unique in that the high temperature requirements force a
temperatures.
heat input as well as an electricity input. SOEC is the least developed
electrolysis method, still in the bench scale. SOEC can be operated at 5.2. Thermochemical water splitting
pressures up to around 3 MPa (Gielen and Simbolotti, 2005).
An oxygen ion is the energy carrier for the steam electrolysis Thermochemical water splitting, as previously described, is one or
reaction. The reactions that take place on the anode and cathode are more endothermic chemical reactions that require heat to split water.
shown in Eqs. (21) and (22) (El-Shafie et al., 2019): The two contenders considered for this study are the sulfur–iodine
Anode: (S–I) and calcium-bromide (Ca–Br) cycles. Thermochemical cycles are
2O− → O2 + 4𝑒− (21) less developed than electrolysis processes. In these cycles, chemicals
2
are recycled throughout the system, so inputs and outputs consist
Cathode: of water and water products, respectively. Though thermochemical
cycles primarily utilize heat, rendering them more cost effective than
H2 O + 2𝑒− → H2 + O− (22)
2 electrolysis processes, they can face problems such as slow response
The most common electrode materials are nickel-zirconia (Ni-YSZ) rates, large size, and involve hazardous chemicals that may produce
cermet at the cathode and strontium-doped lanthanum manganite less pure hydrogen (Boardman, 2017).
(LSM), potentially with YSZ at the anode (O’Brien et al., 2010; Hauch,
2008). The exchange current densities for SOEC are much higher than 5.2.1. Sulfur–iodine cycle (S–I)
previously seen electrodes at 2*10− 1 A/cm2 for the Ni-YSZ cathode and The sulfur–iodine (S–I) cycle is the most developed thermochemical
5.3*10− 1 A/cm2 at the LSM/YSZ anode (Chan and Xia, 2002). Resis- cycle, with a high interest pertaining to hybrid nuclear facilities. It was
tance ranges in literature from 0.25 to 0.4 Ω-cm (O’Brien et al., 2010; proposed by General Atomics in the mid-1970’s (Yildiz and Kazimi,
Bragg-Sitton et al., 2014), but a value of 0.4 Ω-cm was taken to mimic 2006). Two proof of concept facilities and laboratory-scale experiments
have been performed for this cycle (Ping et al., 2018). The S–I cycle
a study done by Idaho National Laboratory (INL), in reference (Bragg-
has been down-selected in the United States as well as other countries,
Sitton et al., 2014). The current density is typically somewhere between
such as Japan, as the most promising thermochemical cycle (El-Emam
alkaline and PEM (0.3–1 A/cm2 (Gielen and Simbolotti, 2005)), and
and Özcan, 2019; Varrin et al., 2011). The chemical reactions that take
around 0.7 A/cm2 for the INL reference study (Bragg-Sitton et al.,
place in each step are shown in the following Ping et al. (2018) and
2014). A large amount of water is required for SOECs, estimated at 83.3
Varrin et al. (2011):
times the hydrogen production, because of cooling (McKellar, 2010).
The process diagram for SOEC is shown in Fig. 8. H2 SO4 𝐷𝑒𝑐𝑜𝑚𝑝𝑜𝑠𝑖𝑡𝑖𝑜𝑛 ∶ 2H2 SO4 → 2SO2 + O2 + 2H2 O (> 800 ◦ C) (23)

10
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

Fig. 9. Sulfur–Iodine (S–I) thermochemical process (Varrin et al., 2011).

Fig. 11. Hybrid Sulfur (HyS) hydrogen production schematic (Varrin et al., 2011).

5.3. Hybrid water splitting

Hybrid water splitting involves a thermochemical cycle in conjunc-


tion with electrolysis. The main contenders in the hybrid splitting
process are the hybrid sulfur cycle developed by Westinghouse and
the copper–chlorine cycle. Hybrid cycles have similar disadvantages
and advantages as thermochemical cycles alone, mainly: slow response
rates, large reactor sizes, utilize hazardous chemicals, and are less
developed than purely electrolysis, but are more cost effective than
electrolysis alone (Boardman, 2017).

Fig. 10. Calcium–Bromine (Ca–Br) also known as the UT-3 thermochemical water 5.3.1. Hybrid sulfur cycle (HyS) or westinghouse cycle
splitting diagram. The hybrid sulfur cycle (HyS) has been the most promising hybrid
Source: Adapted from Sakurai et al. (1996).
cycle, having been down-selected as a candidate for future NHES use in
the United States (Varrin et al., 2011). It was proposed by Westinghouse
in 1975 as a hybrid version of the S–I cycle, and shares the H2 SO4
𝐵𝑢𝑛𝑠𝑒𝑛 ∶ I2 + SO2 + 2H2 O → H2 SO4 + 2HI (120 ◦ C) (24) decomposition step (El-Emam and Özcan, 2019; Yildiz and Kazimi,
HI 𝐷𝑒𝑐𝑜𝑚𝑝𝑜𝑠𝑖𝑡𝑖𝑜𝑛 ∶ 2HI → H2 + I2 (> 300 C) ◦
(25) 2006). In addition, there is an electrolysis step which uses a polymer
based membrane to produce hydrogen (Varrin et al., 2011). The process
In the S–I cycle, the H2 SO4 acts as the oxygen carrier while the HI is shown in Fig. 11.
carries hydrogen (Varrin et al., 2011). Mass balances and optimization The chemical reactions that take place for H2 SO4 decomposition and
studies have been conducted for all reactions to maximize reaction SO2 electrolysis are shown in the following Varrin et al. (2011) and
rates and minimize undesired products from alternative reactions oc- Le Duigou et al. (2007):
curring (Ping et al., 2018). The S–I process can be seen in Fig. 9. Note
that the decomposition reactions are both endothermic, but the Bunsen H2 SO4 𝐷𝑒𝑐𝑜𝑚𝑝𝑜𝑠𝑖𝑡𝑖𝑜𝑛 ∶ 2H2 SO4 → 2SO2 + O2 + 2H2 O (> 800 ◦ C) (30)
reaction is exothermic. The limiting reaction is the H2 SO4 reaction, due 𝐸𝑙𝑒𝑐𝑡𝑟𝑜𝑙𝑦𝑠𝑖𝑠 ∶ SO2 + 2H2 O → H2 SO4 + H2 (< 100 ◦ C) (31)
to the high temperature requirements. This decomposition reaction is
why the S–I process may only be coupled to high temperature nuclear One potential advantage of the HyS cycle exists with the location of
reactors if no external heat is desired. hydrogen production. The processes that require heat must be placed
relatively close to the nuclear plant to minimize losses, which includes
5.2.2. Calcium–bromine cycle (Ca–Br) or UT-3 cycle the decomposition step of this process. However, the electrolysis step,
The calcium–bromine (Ca–Br) cycle was discovered at the Univer- which actually produces the hydrogen, may be placed a safer distance
sity of Tokyo in Japan, which is why it is also referred to as the away from the plant as it only requires electricity (Yildiz and Kazimi,
UT-3 cycle (Yildiz and Kazimi, 2006). While considered a promising 2006).
candidate for nuclear hybrid processes, it is less technologically mature
than its competitor, the S–I cycle, and therefore has not been further se- 5.3.2. Copper–chlorine (Cu–Cl) cycle
lected in down-selection processes (Varrin et al., 2011). The minimum The copper–chlorine cycle, like the Ca–Br cycle, has shown promise
temperature required for this process is around 760 ◦ C (Sakurai et al., but is not as technologically mature as other processes (Varrin et al.,
1996), which potentially could be met with a high temperature reactor, 2011). However, it has a lower minimum temperature requirement,
and also a potential for medium temperature reactors with external around 500 ◦ C (Naterer et al., 2009), than the other thermochemi-
heating. The UT-3 process can be divided into four chemical reactors cal processes which would allow it to couple with medium or high
as shown in Fig. 10. A unique advantage of the UT-3 cycle is that it temperature as opposed to only high temperature reactors.
only involves solids and gases, which are easier to separate than the all There are three Cu–Cl cycles proposed: a 3-step, 4-step, and 5-step
liquid reagents of the S–I cycle (Yildiz and Kazimi, 2006). cycle. The 4-step cycle was chosen for this study because it combines
The chemical reactions and temperatures required that take place two steps of the 5-step cycle to reduce heat requirements and complex-
in each reactor are shown in the following Sakurai et al. (1996) and ity, and has lower heat requirements than the 3-step process (Ozbilen
Teo et al. (2005): et al., 2011). The chemical reactions that take place in the 4-step cycle
Ca 𝑅𝑒𝑎𝑐𝑡𝑜𝑟 1 ∶ CaBr 2 + H2 O → CaO + 2HBr (760 ◦ C) (26) are shown in the following reactions (Naterer et al., 2009), and a

diagram of this process is shown in Fig. 12.
Fe 𝑅𝑒𝑎𝑐𝑡𝑜𝑟 1 ∶ 3FeBr 2 + 4H2 O → Fe3 O4 + H2 + 6HBr (560 C) (27)
◦ HCl 𝑃 𝑟𝑜𝑑𝑢𝑐𝑡𝑖𝑜𝑛 ∶ 2CuCl2 + H2 O → CuO ∗ CuCl2 + 2HCl (400 ◦ C) (32)
Fe 𝑅𝑒𝑎𝑐𝑡𝑜𝑟 2 ∶ Fe3 O4 + 8HBr → 3FeBr 2 + Br 2 + 4H2 O (220 C) (28)
𝐷𝑒𝑐𝑜𝑚𝑝𝑜𝑠𝑖𝑡𝑖𝑜𝑛∕O2 𝑃 𝑟𝑜𝑑𝑢𝑐𝑡𝑖𝑜𝑛 ∶ 2(CuO ∗ CuCl2 ) → 4CuCl + O2 (500 ◦ C)
Ca 𝑅𝑒𝑎𝑐𝑡𝑜𝑟 2 ∶ 2CaO + 2Br 2 → 2CaBr 2 + O2 (572 ◦ C) (29)
(33)

11
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

Fig. 12. Copper–Chlorine (Cu–Cl) hybrid cycle diagram.


Source: Adapted from Naterer et al. (2009).

Fig. 13. Process flow diagram for (a) Alkaline water electrolysis (Holladay et al., 2009;
Rashid et al., 2015; Santos et al., 2013; Phillips et al., 2017; Carmo et al., 2013;
𝐸𝑙𝑒𝑐𝑡𝑟𝑜𝑙𝑦𝑠𝑖𝑠 ∶ 2CuCl + 2HCl → H2 + 2CuCl2 (> 430 ◦ C) (34) Miles et al., 1976; Genovese et al., 2009), and (b) PEM water electrolysis (Holladay
◦ et al., 2009; Rashid et al., 2015; Marangio et al., 2009; Siracusano et al., 2018; Carmo
𝐷𝑟𝑦𝑒𝑟 ∶ CuCl2 (𝑎𝑞) → CuCl2 (𝑠) (< 100 C) (35) et al., 2013; Genovese et al., 2009; Choi et al., 2004) given a nominal low temperature
reactor.
The oxygen production step requires the highest temperatures. The
electrolysis process shown only requires low (ambient) temperatures,
but is combined with another (chemical) step that increases the mini-
mum temperature to around 430 ◦ C.
The Cu–Cl electrolysis step has been experimentally demonstrated
for long durations by Canadian Nuclear Laboratories, Pennsylvania
State University, and the University of Ontario Institute of Technol-
ogy (Naterer et al., 2017). The electrolyzer can be PEM but more
research is needed to develop a membrane that limits Cu crossover
to the cathode (hydrogen producing side) (Naterer et al., 2017). One
challenge with the Cu–Cl process is related to materials due to lim-
ited knowledge of corrosion resistance in materials by molten Cu–Cl
(Naterer et al., 2017).

6. Results and discussion

NHES hydrogen candidates were compared based on their energy


input and hydrogen production rates for low, medium, and high tem-
perature reactors. They were also compared relative to their technology
readiness level (TRL) and cost per kg of H2 production.

6.1. Process flow diagrams

Process flow energy and material input and output diagrams were
created for the three reactor categories. These models assume the entire
reactor thermal output (as shown in Table 2) is converted to a either
electricity, heat, or both. The power cycle conversion rate is assumed to
be 40%. As the reactor temperature category increases, there are more
possible hydrogen production methods.

Fig. 14. Process flow diagram for (a) Alkaline water electrolysis (Holladay et al., 2009;
6.1.1. Low temperature reactors
Rashid et al., 2015; Santos et al., 2013; Phillips et al., 2017; Carmo et al., 2013;
Low temperature reactors cover the category of LWRs, which may Miles et al., 1976; Genovese et al., 2009), (b) PEM water electrolysis (Holladay et al.,
output water at temperatures around 300 ◦ C. The only nuclear hybrid 2009; Rashid et al., 2015; Marangio et al., 2009; Siracusano et al., 2018; Carmo et al.,
2013; Genovese et al., 2009; Choi et al., 2004), (c) SOEC electrolysis (McKellar, 2010;
candidates that operate below this temperature are alkaline and PEM
Rashid et al., 2015; Hauch, 2008; Bragg-Sitton et al., 2014; Chan and Xia, 2002), (d)
electrolysis. The average LWR power output is assumed to be 2500 SMR (McKellar, 2010), and (e) Cu–Cl cycle (Nikolaidis and Poullikkas, 2017; Ozbilen
MWt, and all power is assumed to be used for hydrogen production. et al., 2011) given a nominal medium temperature reactor.
Therefore, the following energy and material process flow diagrams,
Fig. 13, shows electricity and heat (which total to 2500 MWt), and
water consumption to produce hydrogen and oxygen. Given the as- 6.1.2. Medium temperature reactors
sumptions made for modeling alkaline and PEM electrolysis, both Medium temperature reactors encompass AGRs, SFRs, SCWRs, LFRs,
produce reasonably similar amounts of hydrogen for a given thermal and MSFRs, to name a few, with an output temperature less than 700–
power input. 750 ◦ C. The nominal size was calculated to be 1800 MWt (Goldberg

12
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

and Rosner, 2011). All processes that can couple with low temper-
ature reactors (alkaline and PEM electrolysis) can also couple with
medium temperature reactors but are less ideal due to wasted heat.
Additionally, SOEC electrolysis and Cu–Cl cycle can operate at these
reactor temperatures. An energy and material process flow diagram
was also created for nuclear-assisted SMR, which requires 700–1000 ◦ C
steam and the additional input of natural gas (which is combusted to
give the heat requirement). Fig. 14 shows these process flow diagrams.
Note that since SMR only requires a small amount of electricity, an
extremely large amount of hydrogen could be produced with the 720
MWe provided by a nuclear reactor. In reality, the amount of natural
gas used for this much electricity would probably be a limiting factor.
Excluding SMR, SOEC and the Cu–Cl cycles can produce much more
hydrogen than either low temperature electrolysis method.

6.1.3. High temperature reactors


High temperature Generation IV reactors, such as the VHTR, MSR,
or GFR, will likely be able to provide outlet temperatures higher
than 900 ◦ C, and possibly up to 950 ◦ C. The average power output
of the three previously listed reactors hovers around 1400 MWt. All
candidates could work with these reactors, but again the ones that
require higher operating temperatures will be more efficient since less
heat needs to be wasted in the process. However, as shown in Fig. 15,
the Cu–Cl process would produce the most hydrogen, but would waste
heat to lower the temperature. The SMR would still be limited by
natural gas, not electricity.

6.2. Technology readiness level

The technology readiness level (TRL) is a means to describe the ex-


tent of development necessary to reach the state of commercialization.
It can range from theoretical principle, at TRL 1, to operational plant, at
TRL 9. Fig. 16 shows the correspondence of TRL number to technology
status (Boardman, 2017).
The TRL was estimated for all hydrogen production candidates
(Boardman, 2017; El-Emam and Özcan, 2019; Ping et al., 2018; Varrin
et al., 2011; Naterer et al., 2017). The minimum TRL of any component
of the technology was taken, as the limiting factor in development
is the component with the lowest TRL. Fig. 17 shows the TRL levels
for all previously described candidates. SMR and alkaline electrolysis
are the only current commercially available technologies. Electrolysis
processes are also more developed than any thermochemical or hybrid
process.

6.3. Cost comparison

Table 4 includes a range of cost per unit kg of hydrogen for


each method. Thermal-based water splitting methods are expected to
be more cost effective because the cost of electricity is significantly
higher than the cost of heat (due to conversion) (El-Emam and Özcan,
2019). Similarly, production costs for electrolysis, requiring significant
electricity, is highly dependent on electricity prices. Fig. 15. Process flow diagram for (a) Alkaline water electrolysis (Holladay et al., 2009;
As seen in the process flow diagrams, SMR technology incurs ex- Rashid et al., 2015; Santos et al., 2013; Phillips et al., 2017; Carmo et al., 2013;
Miles et al., 1976; Genovese et al., 2009), (b) PEM water electrolysis (Holladay et al.,
tremely high costs due to orders of magnitude more hydrogen pro- 2009; Rashid et al., 2015; Marangio et al., 2009; Siracusano et al., 2018; Carmo et al.,
duced. In reality, SMR is the cheapest technology currently in a $/kg 2013; Genovese et al., 2009; Choi et al., 2004), (c) SOEC electrolysis (McKellar, 2010;
scenario (Table 2). Otherwise SOEC and S–I would likely be the cheap- Rashid et al., 2015; Hauch, 2008; Bragg-Sitton et al., 2014; Chan and Xia, 2002), (d)
est when coupled with a high temperature reactor, as these technolo- SMR (McKellar, 2010), (e) S–I cycle (Varrin et al., 2011), (f) Cu–Cl cycle (Nikolaidis
and Poullikkas, 2017; Ozbilen et al., 2011), (g) Ca–Br cycle (Sakurai et al., 1996; Teo
gies have previously been chosen for further development (Varrin et al.,
et al., 2005), and (h) HyS cycle (Varrin et al., 2011) given a nominal high temperature
2011). There are situations where hybrid systems may be profitable — reactor.
when hydrogen, electricity, and natural gas prices are high, or severe
CO2 costs are included (Ruth et al., 2017). However, with the volatility
of prices, flexible hybrid configurations are not currently valuable nuclear power plants to add electrolysis-based hydrogen production
enough to overcome capital costs (Ruth et al., 2017). Further policy will
to existing LWR plants, to demonstrate the capabilities and long-term
likely need to be implemented (aiming at cleaner energy production)
to make hybrid systems profitable (Ruth et al., 2017). The Department competitiveness of nuclear hydrogen production and improve plant
of Energy (DOE), as of late 2019, has issued contracts with three economics (Patel, 2019).

13
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

Table 4
Inputs and outputs per kg of hydrogen produced, TRL, and unit cost for candidate NHES technologies (Holladay et al., 2009; Nikolaidis and Poullikkas, 2017; Boardman, 2017;
El-Emam and Özcan, 2019; McKellar, 2010; Rashid et al., 2015; Santos et al., 2013; Phillips et al., 2017; Siracusano et al., 2018; Bragg-Sitton et al., 2014; Miles et al., 1976;
Genovese et al., 2009; Choi et al., 2004; Chan and Xia, 2002; Ping et al., 2018; Varrin et al., 2011; Sakurai et al., 1996; Teo et al., 2005; Ozbilen et al., 2011; Naterer et al.,
2017; Ruth et al., 2017; James et al., 2016; Spath and Mann, 2000).
Electrolysis Thermochemical Hybrid
Alkaline PEM SOEC SMR S–I Ca–Br HyS Cu–Cl
T𝑀𝑎𝑥 (C) 60 80 30 60 90 700 800 900 870 710 910 760 710 910 500
Pressure (MPa) 0.1 0.1 0.157 4.1 3.85 2 4 0.1–2
TRL 9 6–8 5 9 4 <3 3–4 <3
H2 Yield efficiency (HHV) (%)a 29.8 29.4 25.2 27.0 27.0 36.0 35.8 35.6 79.4 35.6b 25.2b 39.0 31b 35.6b 43.2
Inputs (/kg H2 )
Electricity (MJ) 180.0 181.9 214.9 199.5 199.1 146.7 146.4 146.2 1.4 51.5 75.0 25.3 59.0 82.8 67.2
Heat (MJ) 26.2 27.4 24.4 26.2 28.1 26.7 29.7 32.7 – 269.3 375.0 301.4 309.8 178.0 161.1
Water (kg) 11.5 11.5 11.5 11.5 11.5 83.3 83.3 83.3 10.3 9.4 9.0 9.0 9.0 9.0 9.0
Nat gas (kg) – – – – – – – – 2.9 – – – – – –
Outputs (/kg H2 )
H2 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1
O2 8 8 8 8 8 8 8 8 0 8 8 8 8 8 8
CO2 0 0 0 0 0 0 0 0 4.7f 0 0 0 0 0 0
Production cost (est. $US2019 ) $5.92c $3.56–5.46c $2.24–3.73c $1.54–2.30d $2.18–5.65c $7.06e $2.29–6.27c $2.36–3.86c
a Assuming a power cycle conversion efficiency of 40%.
b Note efficiency highly dependent on process assumptions, which change between these studies, and if temperature is lower than minimum required, external heat must be supplied
(710 ◦ C) studies (Varrin et al., 2011).
c Cost based on US$2019 (El-Emam and Özcan, 2019).

d
Cost based on US$2017 and incorporating inflation to 2019 (Ruth et al., 2017).
e
Cost based on US$1995 and incorporating inflation to 2019 (Sakurai et al., 1996).
f
CO2 estimate may be low; alternative literature quotes as much as 10.6 kg/kg H2 (Spath and Mann, 2000).

Fig. 16. TRL number descriptions (Boardman, 2017).

7. Conclusion

With increasing renewable energy sources along with increasing


pressure to promote clean, sustainable energy, nuclear hybrid energy
systems are a potential way to keep nuclear competitive. With renew-
able energy, grid power becomes increasingly volatile, leading to less
base-load power systems, like nuclear, to flexible energy producers.
These flexible energy producers are typically greenhouse gas emitting
coal or natural gas plants. With nuclear hydrogen hybrid technology,
an alternative commodity, hydrogen, could be produced to allow for
grid load-following. This hydrogen can be used for energy storage
and convert back to energy during high electricity prices, or has the
potential to be an energy carrier, to be used as fuel for transportation.
Hydrogen does not produce CO2 , rendering it a much cleaner fuel than
currently used petroleum or natural gas.
Fig. 17. TRL of various NHES hydrogen production candidates (Boardman, 2017; Several hydrogen producing nuclear hybrid technologies were in-
El-Emam and Özcan, 2019; Ping et al., 2018; Varrin et al., 2011; Naterer et al., 2017). vestigated in this study. The NHES candidate technologies (along with
the comparison SMR), including temperature, pressure, TRL, efficiency,
hydrogen cost, and inputs and outputs are summarized in Table 4.
These technologies only require water, heat and/or electricity, all of

14
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

which are abundant around a nuclear plant, to produce hydrogen. Al- Grimes, C., Varghese, O., Ranjan, S., 2007. Light, Water, Hydrogen: The Solar
kaline water electrolysis is currently the only widespread, commercially Generation of Hydrogen by Water Photoelectrolysis. Springer Science & Business
Media.
available, large scale clean hydrogen production technology, but other
Guenot, B., Cretin, M., Lamy, C., 2015. Clean hydrogen generation from the electro-
technologies have shown rapid development and promise, along with catalytic oxidation of methanol inside a proton exchange membrane electrolysis
potentially lower production costs. cell (PEMEC): effect of methanol concentration and working temperature. J. Appl.
Additionally, these hydrogen production technologies operate at Electrochem. 45 (9), 973–981.
Hauch, A., 2008. Solid Oxide Electrolysis Cells: Performance and Durability. Risø
various temperatures, which allows them to match closely with the
National Laboratory.
current (Generation III) and future (Generation IV) reactor fleet. These Holladay, J.D., Hu, J., King, D.L., Wang, Y., 2009. An overview of hydrogen production
nuclear reactors may be loosely subdivided into three categories: low technologies. Catal. Today 139 (4), 244–260.
output temperatures (<300 ◦ C), medium (<750 ◦ C), and high tem- James, B.D., DeSantis, D.A., Saur, G., 2016. Hydrogen Production Pathways Cost
peratures (<950 ◦ C). Within these categories, only low temperature Analysis (2013–2016). Technical Report, Strategic Analysis Inc., Arlington, VA
(United States), No. DOE-StrategicAnalysis-6231-1.
electrolysis is currently available for low temperature reactors, but
Kabza, A., 2010. Another fuel cell formulary.
thermochemical and electrolysis methods could potentially be coupled Le Duigou, A., Borgard, J.-M., Larousse, B., Doizi, D., Allen, R., Ewan, B.C., Priest-
with high temperature systems. Energy and material process flow di- man, G.H., Elder, R., Devonshire, R., Ramos, V., et al., 2007. HYTHEC: An EC
agrams were created for hybrid technologies at all reactor category funded search for a long term massive hydrogen production route using solar and
nuclear technologies. Int. J. Hydrogen Energy 32 (10–11), 1516–1529.
levels. When coupled with around 2000 MWt power from a nuclear
Marangio, F., Santarelli, M., Cali, M., 2009. Theoretical model and experimental
reactor, these technologies can generate between 200–500 tons/day of analysis of a high pressure PEM water electrolyser for hydrogen production. Int.
hydrogen, with no CO2 emissions. Current hybrid technologies are not J. Hydrogen Energy 34 (3), 1143–1158.
cost effective, and will require either increased electricity prices, CO2 McKellar, M., 2010. Nuclear-Integrated Hydrogen Production Analysis. Technical Re-
subsidies, or increased hydrogen prices to overcome capital costs. How- port, Idaho National Lab.(INL), Idaho Falls, ID (United States), TEV-693, Rev
1.
ever, new contracts from DOE will provide funding to nuclear power Miles, M., Kissel, G., Lu, P., Srinivasan, S., 1976. Effect of temperature on electrode ki-
plants to demonstrate hydrogen production capabilities to improve the netic parameters for hydrogen and oxygen evolution reactions on nickel electrodes
competitiveness of hybrid technology and improve plant economics. in alkaline solutions. J. Electrochem. Soc. 123 (3), 332–336.
Naterer, G., Suppiah, S., Lewis, M., Gabriel, K., Dincer, I., Rosen, M.A., Fowler, M.,
Rizvi, G., Easton, E., Ikeda, B., et al., 2009. Recent Canadian advances in nuclear-
Declaration of competing interest based hydrogen production and the thermochemical Cu–Cl cycle. Int. J. Hydrogen
Energy 34 (7), 2901–2917.
The authors declare that they have no known competing finan- Naterer, G., Suppiah, S., Rosen, M., Gabriel, K., Dincer, I., Jianu, O., Wang, Z.,
Easton, E., Ikeda, B., Rizvi, G., et al., 2017. Advances in unit operations and
cial interests or personal relationships that could have appeared to
materials for the CuCl cycle of hydrogen production. Int. J. Hydrogen Energy 42
influence the work reported in this paper. (24), 15708–15723.
Nikolaidis, P., Poullikkas, A., 2017. A comparative overview of hydrogen production
References processes. Renew. Sustain. Energy Rev. 67, 597–611.
O’Brien, J., Stoots, C., Herring, J., McKellar, M., Harvego, E., Sohal, M., Condie, K.,
2010. High Temperature Electrolysis for Hydrogen Production from Nuclear
Antkowiak, M., Boardman, R., Bragg-Sitton, S., Cherry, R., Ruth, M., 2012. Summary
Energy–TechnologySummary. Technical Report, Idaho National Laboratory (INL),
Report of the INL-JISEA Workshop on Nuclear Hybrud Energy Systems. Technical
No. INL/EXT-09-16140.
Report, Idaho National Laboratory (INL).
OnSite, P., 2019. Proton onsite introduces world’s first PEM megawatt electrolyzer for
Boardman, R.D., 2017. Figures of Merit for Technical and Economic Assessment
the growing global energy storage market. https://www.protononsite.com/news-
of Nuclear Hydrogen Hybrid Energy Systems. Technical Report, Idaho National
events/proton-onsite-introduces-worlds-first-pem-megawatt-electrolyzer-growing-
Lab.(INL), Idaho Falls, ID (United States), No. INL/EXT-17-41484.
global-energy. (Accessed 18 July 2019).
Bossel, U., Eliasson, B., 2003. Energy and the Hydrogen Economy. Methanol Institute,
Ozbilen, A., Dincer, I., Rosen, M., 2011. Environmental evaluation of hydrogen
Arlington, VA.
production via thermochemical water splitting using the Cu–Cl cycle: a parametric
Bragg-Sitton, S.M., Boardman, R.D., Cherry, R.S., Deason, W.R., McKellar, M.G.,
study. Int. J. Hydrogen Energy 36 (16), 9514–9528.
2014. An Analysis of Methanol and Hydrogen Production via High-Temperature
Patel, S., 2019. Three more nuclear plant owners will demonstrate hydrogen
Electrolysis Using the Sodium Cooled Advanced Fast Reactor. Technical Report,
production. Power https://www.powermag.com/three-more-nuclear-plant-owners-
Idaho National Laboratory (INL).
will-demonstrate-hydrogen-production/. (Accessed 5 January 2020).
Carmo, M., Fritz, D.L., Mergel, J., Stolten, D., 2013. A comprehensive review on PEM Phillips, R., Edwards, A., Rome, B., Jones, D.R., Dunnill, C.W., 2017. Minimising the
water electrolysis. Int. J. Hydrogen Energy 38 (12), 4901–4934. ohmic resistance of an alkaline electrolysis cell through effective cell design. Int.
Chan, S., Xia, Z., 2002. Polarization effects in electrolyte/electrode-supported solid J. Hydrogen Energy 42 (38), 23986–23994.
oxide fuel cells. J. Appl. Electrochem. 32 (3), 339–347. Ping, Z., Laijun, W., Songzhe, C., Jingming, X., 2018. Progress of nuclear hydrogen
Choi, P., Bessarabov, D.G., Datta, R., 2004. A simple model for solid polymer electrolyte production through the iodine–sulfur process in China. Renew. Sustain. Energy
(SPE) water electrolysis. Solid State Ion. 175 (1–4), 535–539. Rev. 81, 1802–1812.
Dagle, R.A., Dagle, V., Bearden, M.D., Holladay, J.D., Krause, T.R., Ahmed, S., 2017. An Rand, D.A.J., Dell, R.M., 2007. Hydrogen Energy: Challenges and Prospects. Royal
Overview of Natural Gas Conversion Technologies for Co-Production of Hydrogen Society of Chemistry.
and Value-Added Solid Carbon Products. Technical Report, Pacific Northwest Rashid, M.M., Al Mesfer, M.K., Naseem, H., Danish, M., et al., 2015. Hydrogen
National Lab.(PNNL), Richland, WA (United States); Argonne . . . , No. PNNL-26726; production by water electrolysis: a review of alkaline water electrolysis, PEM water
ANL-17/11. electrolysis and high temperature water electrolysis. Int. J. Eng. Adv. Technol. 4
Department of Energy (DOE) Office of Energy Efficiency & Renewable Energy, (3), 2249–8958.
2019. Hydrogen production. https://www.energy.gov/eere/fuelcells/hydrogen- Rosen, M.A., 2010. Advances in hydrogen production by thermochemical water
production. (Accessed 14 June 2019). decomposition: a review. Energy 35 (2), 1068–1076.
El-Emam, R.S., Özcan, H., 2019. Comprehensive review on the techno-economics of Ruth, M., Cutler, D., Flores-Espino, F., Stark, G., 2017. The Economic Potential
sustainable large-scale clean hydrogen production. J. Cleaner Prod. 220, 593–609. of Nuclear-Renewable Hybrid Energy Systems Producing Hydrogen. Technical
El-Shafie, M., Kambara, S., Hayakawa, Y., 2019. Hydrogen production technologies Report, National Renewable Energy Lab.(NREL), Golden, CO (United States), No.
overview. J. Power Energy Eng. 7, 107–154. NREL/TP-6A50-66764.
Forsberg, C., 2019. Meeting Low-Carbon Industrial Heat Demand with High- Sakurai, M., Bilgen, E., Tsutsumi, A., Yoshida, K., 1996. Adiabatic UT-3 thermochemical
Temperature Reactors Using Co-Generation and Heat Storage, Vol. 120. process for hydrogen production. Int. J. Hydrogen Energy 21 (10), 865–870.
Transactions of the American Nuclear Society. Santos, D.M., Sequeira, C.A., Figueiredo, J.L., 2013. Hydrogen production by alkaline
Genovese, J., Harg, K., Paster, M., Turner, J., 2009. Current (2009) State-of-the-Art water electrolysis. Química Nova 36 (8), 1176–1193.
Hydrogen Production Cost Estimate Using Water Electrolysis: Tndependent Review. Sapountzi, F.M., Gracia, J.M., Fredriksson, H.O., Niemantsverdriet, J.H., et al., 2017.
Technical Report, National Renewable Energy Laboratory (NREL). Electrocatalysts for the generation of hydrogen, oxygen and synthesis gas. Prog.
Gielen, D., Simbolotti, G., 2005. Prospects for Hydrogen and Fuel Cells. International Energy Combust. Sci. 58, 1–35.
Energy Agency (IEA). Schmidt, O., Gambhir, A., Staffell, I., Hawkes, A., Nelson, J., Few, S., 2017. Future cost
Goldberg, S.M., Rosner, R., 2011. Nuclear Reactors: Generation to Generation. American and performance of water electrolysis: An expert elicitation study. Int. J. Hydrogen
Academy of Arts and Sciences, Cambridge, USA. Energy 42 (52), 30470–30492.

15
R. Pinsky et al. Progress in Nuclear Energy 123 (2020) 103317

Siracusano, S., Trocino, S., Briguglio, N., Baglio, V., Aricò, A., 2018. Electrochemical Varrin, R., Reifsneider, K., Scott, D., Irving, P., Rolfson, G., 2011. NGNP Hydrogen
impedance spectroscopy as a diagnostic tool in polymer electrolyte membrane Technology Down Selection; Results of the Independent Review Team Evaluation.
electrolysis. Materials 11 (8), 1368. Dominion Engineering report for Idaho National Laboratory.
Spath, P.L., Mann, M.K., 2000. Life Cycle Assessment of Hydrogen Production Via Yildiz, B., Kazimi, M.S., 2006. Efficiency of hydrogen production systems using
Natural Gas Steam Reforming. Technical Report, National Renewable Energy Lab., alternative nuclear energy technologies. Int. J. Hydrogen Energy 31 (1), 77–92.
Golden, CO (US), No. NREL/TP-570-27637. Zeng, K., Zhang, D., 2010. Recent progress in alkaline water electrolysis for hydrogen
Teo, E., Brandon, N., Vos, E., Kramer, G., 2005. A critical pathway energy efficiency production and applications. Prog. Energy Combust. Sci. 36 (3), 307–326.
analysis of the thermochemical UT-3 cycle. Int. J. Hydrogen Energy 30 (5),
559–564.

16

You might also like