You are on page 1of 54

3.

2 Ammonia Production
Ibrahim Dincer and Yusuf Bicer, University of Ontario Institute of Technology, Oshawa, ON, Canada
r 2018 Elsevier Inc. All rights reserved.

3.2.1 Introduction 42
3.2.2 Fundamentals: Ammonia Production Methods 44
3.2.2.1 Hydrocarbon Cracking and Reforming 47
3.2.2.1.1 Steam methane reforming-based ammonia production 47
3.2.2.1.2 Partial oxidation of heavy oil-based ammonia production 48
3.2.2.1.3 Methane dissociation for ammonia production 49
3.2.2.2 Water Electrolysis 49
3.2.2.2.1 Wind electrolysis-based ammonia production 49
3.2.2.2.2 Geothermal electrolysis-based ammonia production 50
3.2.2.2.3 Solar energy electrolysis-based ammonia production 50
3.2.2.2.4 Nuclear-based ammonia production 51
3.2.2.3 Gasification 52
3.2.2.3.1 Coal gasification-based ammonia production 53
3.2.2.3.2 Biomass gasification-based ammonia production 53
3.2.2.4 Novel Ammonia Production Methods 53
3.2.2.4.1 Liquid electrolyte-based systems 60
3.2.2.4.2 Composite membrane-based systems 60
3.2.2.4.3 Solid state electrolyte 61
3.2.2.4.4 Ceramic/inorganic proton conducting solid electrolyte-based systems 61
3.2.2.4.5 Polymer membrane-based systems 61
3.2.2.4.6 O2  conducting membrane materials and ammonia synthesis systems 62
3.2.3 Case Study 1: Electrochemical Ammonia Production 62
3.2.3.1 Electrochemical Ammonia Synthesis 63
3.2.3.2 Case Study 1 Results and Discussion 65
3.2.3.3 Case Study 1 Conclusions 66
3.2.4 Case Study 2: Life Cycle Assessment of Various Fuels Including Ammonia 67
3.2.4.1 Life Cycle Assessment Steps 68
3.2.4.1.1 Goal and scope definition 68
3.2.4.1.2 Inventory analysis 68
3.2.4.1.3 Impact assessment 68
3.2.4.1.4 Interpretation of results and improvement 69
3.2.4.1.4.1 Depletion of abiotic resources 69
3.2.4.1.4.2 Human toxicity 69
3.2.4.1.4.3 Ozone depletion 69
3.2.4.1.4.4 Global warming 69
3.2.4.1.4.5 Human health 69
3.2.4.1.4.6 Ecosystem quality 70
3.2.4.1.4.7 Resources 70
3.2.4.2 Case Study 2 Results and Discussion 70
3.2.4.3 Case Study 2 Conclusions 74
3.2.5 Case Study 3: Comparative Ammonia Production, Storage and Transport Scenarios 74
3.2.5.1 Case 1: Ammonia Production via High-Pressure Electrolysis 74
3.2.5.2 Case 2: Ammonia Production via Liquid Nitrogen and High-Pressure Electrolysis 75
3.2.5.3 Ammonia Fueled Generators for Stand-Alone Power Production by Transporting Ammonia via Tanker
Trucks 77
3.2.5.4 Renewable Energy-Based On-Site Ammonia Production and Utilization in Ammonia Fueled Generators for
Stand-Alone Power Production 78
3.2.5.5 Cost Comparisons 78
3.2.5.6 Case Study 3 Conclusions 80
3.2.6 Case Study 4: Ammonia Production Scenarios in Canada 80
3.2.6.1 Ammonia Production Using Water Electrolysis From Low-Cost Hydropower and Wind Energy 80
3.2.6.2 Ammonia Production From Steam Methane Reforming With CO2 Capture and Sequestration 81
3.2.6.2.1 Current ammonia retail prices 82

Comprehensive Energy Systems, Volume 3 doi:10.1016/B978-0-12-809597-3.00305-9 41


42 Ammonia Production

3.2.6.3 Cost Analyses Results 82


3.2.6.3.1 Storage cost 83
3.2.6.3.2 Transportation cost 85
3.2.6.3.3 Ammonia production in Ontario 86
3.2.6.3.4 Ammonia production in Newfoundland and Labrador 87
3.2.6.3.5 Ammonia production in Alberta 88
3.2.6.3.6 On-site ammonia production in Northwestern Ontario using renewable resources 89
3.2.6.4 Case Study 4 Conclusions 89
3.2.7 Future Directions 90
3.2.8 Closing Remarks 91
Acknowledgment 91
References 91
Further Reading 92
Relevant Websites 93

Nomenclature M Molarity (m)


e Charge of electron (1.60217657  1019 C) n_ Mol flow rate (mol/s)
F Faraday constant (c/mol) Q_ Heat rate (W)
h Enthalpy (kJ/kg) t Time (s)
i Current density (A/m2) V Volume (L)
m_ Mass flow rate (g/s) W _ Work rate (W)

Greek Letters Z Efficiency

Acronyms LCA Life cycle assessment


CAD Canadian dollar LCI Life cycle inventory
CCS Carbon capture storage LHV Lower heating value
CFBG Circulating fluidized bed gasification LNG Liquefied natural gas
CML Center of Environmental Science of Leiden PEM Polymer electrolyte membrane
University SCR Selective catalytic reduction
DOE Department of energy SDC Samaria-doped ceria
GHG Greenhouse gas SFCN Sr2Fe1-xCoxNbO6-δ
GTA Greater Toronto area SMR Steam methane reforming
GWP Global warming potential SSAS Solid state ammonia synthesis
HB Haber–Bosch SSPC Solid-state proton conductors
HHV Higher heating value TKM Ton-kilometer
HTP Human toxicity potentials USES Uniform system for the evaluation of
IPCC Intergovernmental panel on climate change substances
ISO International Organization for Standardization WMO World Meteorological Organization

Subscripts elec Electrical


el Electricity therm Thermal

3.2.1 Introduction

In the world, about 72% of ammonia is produced by steam methane reforming (SMR) from natural gas (Fig. 1). In terms
of conventional resources, naphtha, heavy fuel oil (HFO), coal, natural gas, coke oven gas, and refinery gas can be used as
feedstock for ammonia production. In China, coal is intensively used as a main source of energy, and is generally characterized
by high energy intensities. Natural gas costs represent about 70–90% of the manufacturing cost of ammonia. Subsequently, since
ammonia production is based on natural gas in the SMR method, if natural gas prices rise, production costs for ammonia increase
in parallel [1,2].
Ammonia Production 43

Naphta Others
1% 1%
Fuel oil
4%

Coal
22%

Natural gas
72%

Fig. 1 Sources of ammonia production based on feedstock use. Data from Industrial Efficiency Technology Database. Ammonia| Measures.
Available From: http://ietd.iipnetwork.org/content/ammonia#key-data; 2017 [accessed 07.01.17]; International Energy Agency. Energy technology
perspectives pathways to a clean energy system 2012; France.

45,000
42,290
40,000
Ammonia production (million tonnes)

35,000

30,000

25,000

20,000

15,000
11,200 10,441
10,000
7704
5100 4600 4200 4000
5000
2600 2363
0
China India Russian United Trinidad Indonesia Ukraine Canada Saudi Germany
Federation states and Arabia
Tobago
Fig. 2 Worldwide ammonia production amounts by country. Data from United States Geological Survey (USGS) Minerals Resources Program.
USGS|science for a changing world. Available From: https://www.usgs.gov/; 2017 [accessed 07.01.17].

Ammonia gas is very soluble in water. At ambient temperature and atmospheric pressure, ammonia is an alkaline, colorless gas
with a pungent and suffocating odor. Substantial storage and delivery systems are already available for NH3 used in fertilizer
applications. NH3 has been transported by ship, barge, rail car, truck, and pipeline for decades. It is recognized among the top
three chemicals transported annually. Large capacities which correspond to 20,000–30,000 t and low-cost NH3 storage tanks are
currently installed in many parts of the United States [3] whereas China appears to be the top producer as shown in Fig. 2.
One of the mostly used transportation method for ammonia is pipelines. Therefore long distance transport of ammonia is
classically held by using pipelines [4]. A total of 4830 km carbon steel pipeline network is already employed in the United States to
transport ammonia from port and production facilities to agricultural areas for use as a fertilizer. There are currently storage
services and terminals located along the pipeline to support operations. In Iowa, United States, there is more than 800 ammonia
retail places [3].
It is significant to note that pressurized NH3 storing and distribution substructures are very similar in design and performance
to propane distribution infrastructure since they are both compressed fluids at moderate pressure. The important and confirmed,
worldwide availability of propane-fueled cars and heaters, presents a familiar example of how ammonia fuel systems would work
physically. An ammonia pipeline from the Gulf of Mexico to Minnesota and with branches to Ohio and Texas has served the NH3
manufacturing industry for several periods as shown in Fig. 3. Since ammonia can be sent and stored in mild steel pipelines, any
natural gas or petroleum pipeline could be cost-effectively rehabilitated to transport NH3. Presently, there are nearly two million
44 Ammonia Production

Up
pe
South Dakota Minnesota

r Mi
ss
Wisconsin

iss
ipp
Michigan

iR
iver
Nebraska

Ohio
Iowa

r
ve
Ri
M

Illinois
is s Illinois
ou Indiana
ri R
Kansas i ve r
e

r
Riv
O hi o

Missouri Kentucky

Tennessee

er
Oklahoma

i Riv
Arkansas

sipp
Lower Missis

Alabama
Mississippi

Texas

Magellan Midstream Partners, LP Louisiana


NuStar Energy LP
Terminal

Fig. 3 Ammonia transportation pipelines in United States. Modified from Comparisons. NH3 fuel association. Available From: https://
nh3fuelassociation.org/comparisons/; 2017 [accessed 07.01.17].

miles of natural gas pipeline in the United States. This pipelines could be transformed to transmit NH3, making NH3 fuel readily
available to almost every community [5].
Though there are numerous methods for ammonia synthesis, usually two different ammonia synthesis techniques are available
in the world, namely, Haber–Bosch process and electrochemical ammonia synthesis process as exemplified in Fig. 4. Solid state
ammonia synthesis (SSAS) is one of the mostly known methods among electrochemical options. In both methods, nitrogen is
supplied through air separation system. Cryogenic air separation is the most efficient and cost-effective technology for producing
large amount of oxygen, nitrogen, and argon [6]. Cryogenic technology can also produce high-purity nitrogen as a valuable by-
product stream at comparatively low incremental cost. Among other air separation processes, cryogenic air separation has highly
developed and advanced technology. If ammonia is produced in high quantities, the required nitrogen could be produced at a
low-cost and high efficient way using cryogenic air separation. However, in the decentralized ammonia production, the production
capacities would be lower than central ammonia production plants. In this case, the cost of nitrogen production may also play
important role in total ammonia cost. Therefore, novel methods of nitrogen production at smaller scale and lower cost need to be
investigated. The required power might be supplied either from conventional or alternative sources as shown in Fig. 4.

3.2.2 Fundamentals: Ammonia Production Methods

The Haber–Bosch process is the most common method to produce ammonia [7]. It is an exothermic process that combines
hydrogen and nitrogen in 3:1 ratio to produce ammonia as illustrated in Fig. 4. The reaction is facilitated by catalyst and the
optimal temperature range is 450–6001C. In Haber–Bosch process, the feasibility of ammonia production basically depends on
Ammonia Production 45

Ammonia (NH3)

Hydrogen (H2) Nitrogen (N2)

Hydrocarbon
Water
cracking and Gasification Cryogenic Non-cryogenic
electrolysis
reforming

(A)

CO2 Water

Natural
Conventional

gas H2
Steam methane
reforming

N2

Water
Conventional

Haber−Bosch process
Electricity

NH3

H2
Electricity Electrolyzer
Renewable

Air
Air separation unit
Conventional

N2

NH3
Solid state ammonia synthesis (SSAS)
Electricity
Renewable

Water
O2 O2
(B)

Fig. 4 (A) A simple categorization of ammonia production, and (B) a comprehensive classification of ammonia production methods via
conventional and renewable sources.

the methods used to produce hydrogen and nitrogen. The nitrogen and hydrogen gas mixture are compressed to 120–220 bar,
depending on the particular plant before it enters the ammonia synthesis loop [8]. Only a portion of the mixture gas is converted
to ammonia in a single pass through the converter due to the thermodynamic equilibrium of the ammonia synthesis reaction. The
residual gas which is not reacted is passed through the converter multiple times, creating a flow loop for the unreacted gas.
The ammonia in gas form and unconverted mixture gas then arrive at the ammonia recovery section of the synthesis loop. Using
the refrigeration coolers, the temperature of the gasses is decreased from about  10 to about  251C which condenses ammonia
out of the mixture and leave unreacted synthetic gas behind [9].
46 Ammonia Production

Water Water
Pump Boiler
Heat recovery

Oxygen/
water
Recycled
Oxygen/ Superheated
water
water steam

Furnace
Oxygen Flash tank SSAS tube module

Heat
recovery

N2
Compressor

NH3
NH3/N2 (liquid)

Flash tank
N2

Recycled N2

Fig. 5 Ammonia production via solid state ammonia synthesis (SSAS). Adapted from Ganley JC, Holbrook JH, McKinley DE. Solid state ammonia
synthesis. Ammon Fuel Netw Conf 2007.

One of the world’s main challenges is the population increase. Increasing population will obviously require more foods in
addition to the biomass requirement as energy supply. It is a known fact that this will require more fertilizers in agriculture. If the
production of ammonia and other fertilizers continue to be dependent on fossil fuels, the earth will become a place with polluted
air, increased human health risks, decreased biodiversity and increased greenhouse gas (GHG) [10]. A conversion of nitrogen and
hydrogen into ammonia (with a conversion efficiency of 10–15% in single pass) is thermodynamically not very efficient, due to
the high-temperature requirement and other process complications of Haber–Bosch process [11]. Also, environmental con-
tamination is severe, and energy consumption is high because of fossil fuel usage [12,13]. The way of improving Haber–Bosch is
mostly via modifications in the catalyst and heat recovery. Ruthenium-based catalyst instead of an iron-based catalyst is one of the
catalytic improvements which have recently started to be used. In this way, better catalyst enables more ammonia to be formed per
pass over the converter at lower temperatures and pressures which also bring lower energy consumption.
The other developing ammonia production method is SSAS as shown in Fig. 5. This type of ammonia production system
uses a solid state electrochemical process to produce ammonia from nitrogen, water, and electricity. SSAS consumes less
energy and yields higher efficiencies. The required electricity for SSAS process is 7000–8000 kWh/t-NH3, whereas it is 12,000
kWh/t-NH3 for the combination of the electrolyzer and Haber–Bosch synloop [14]. The capital cost may approximately be
calculated as $200,000/t-day-NH3. Compared with electrolyzer combined to a Haber–Bosch synloop which is around
$750,000/t-day-NH3, it is less [14]. In SSAS, a proton conducting membrane is heated to around 5501C. Under same
pressures, nitrogen and water vapor is supplied to each side of the membrane to initiate the reaction. The water vapor
separates into protons and oxygen. By applying an external voltage, the protons are transferred through the membrane, and on
the nitrogen side of the membrane, NH3 is being formed as a result of nitrogen and protons reaction. Since the energy
consumption of the SSAS process is lower, it is evaluated that it will enable producing ammonia at a lower cost than the
Haber–Bosch process. On the other hand, if renewable electricity is used, it does not consume fossil fuel which brings
significant environmental advantages. Since the electrolyzer and Haber–Bosch synloop are eliminated when SSAS system is
used, the SSAS technology could be suitable for renewable energy sources which result in many energetic and financial
advantages [14].
For the Haber–Bosch process, production of ammonia is based on various hydrogen production techniques as shown in Fig. 6.
In contrast, in SSAS process, it relies on generating superheated steam as illustrated in Fig. 7.
Ammonia Production 47

Renewable Conventional Cryogenic Non-cryogenic

Tidal and Ocean Coal


Solar Wind Hydro Biomass Geothermal UCG SMR Nuclear Heavy oil
wave thermal gasification

H2 N2

NH3 synthesis (Haber−Bosch)

Fig. 6 Ammonia production and usage routes by the Haber–Bosch process.

Renewable Conventional Cryogenic Non-cryogenic

Tidal and Ocean Coal


Solar Wind Hydro Biomass Geothermal UCG SMR Nuclear Heavy oil
wave thermal gasification

H2O (Superheated Steam) N2

NH3 (Solid state ammonia synthesis)

Fig. 7 Ammonia production and usage routes through solid state ammonia synthesis (SSAS) process. SMR, steam methane reforming; UCG,
underground coal gasification.

3.2.2.1 Hydrocarbon Cracking and Reforming


In this method, the source of hydrogen required for ammonia synthesis is hydrocarbons mainly consisting of natural gas, heavy
oil, and naphtha.

3.2.2.1.1 Steam methane reforming-based ammonia production


Most hydrogen produced today in the world is made via SMR, a mature production process in which high-temperature
steam (700–10001C) is used to produce hydrogen from a methane source, such as natural gas. In SMR, methane reacts with steam
under 3–25 bar pressure in the presence of a catalyst to produce hydrogen, carbon monoxide, and a relatively small amount of
carbon dioxide. Steam reforming is endothermic – that is, heat must be supplied to the process for the reaction to proceed as
illustrated in Figs. 8 and 9. In the steam reforming processes, steam is taken from the plant steam system, usually from an
extraction turbine. The net consumption according to the stoichiometric conversion is 0.6–0.7 kg/kg NH3, the total supply at S/C
ratio of 3.0 will be about 1.5 kg/kg NH3. In partial oxidation much less steam is fed to the gasification reactor, but additional
steam is needed in shift conversion (1.2 kg/kg NH3 in total). The typical feedstock requirements for modern plants are
48 Ammonia Production

Air
Natural gas

Hydrogeneration

ZnO bed
Cryogenic air Electricity
O2
separation

Hydrogen product slip stream


Steam Catalytic
steam
Natural gas
reforming
fuel

Shift
reaction

Pressure swing Electricity


adsorption

H2 Ammonia N2
synthesis

NH3

Ammonia
storage

NH3

Ammonia usage

Fig. 8 Ammonia production via steam methane reforming.

Nitrogen (N2)

Steam
Steam methane reforming
Electricity
Ammonia synthesis
Natural gas

Ammonia (NH3)

Fig. 9 Energy and material flows in steam methane reforming-based ammonia production.

approximately 22.1 GJ LHV/t NH3. The main energy and material flows in SMR-based ammonia production process are illustrated
in Fig. 9.

3.2.2.1.2 Partial oxidation of heavy oil-based ammonia production


HFO is the residue of crude oil distillation that still flows, waste oil from other industries are often added. It is the fuel used in large
marine vessels because of price advantage. A typical HFO is IF-300 (Intermediate Fuel), which has a viscosity of 300  106 m2/s at
501C (300 cSt), 25  106 m2/s at 1001C, r¼ 990 kg/m3 at 151C, higher heating value (HHV) ¼ 43 MJ/kg, and the flash-point at
60–801C. HFO may have a composition of 88 wt% C, 10 wt% H, 1 wt% S, 0.5 wt% H2O, 0.1 wt% ash, and may contain dispersed
Ammonia Production 49

Heat Water O2

Nitrogen (N2)
Heavy oil partial oxidation
Electricity
Ammonia synthesis
Heavy oil

Ammonia
(NH3)

Fig. 10 Energy and material flows of partial oxidation of heavy oil-based ammonia production.

solid or semisolid particles (asphaltenes, minerals and other leftovers from the oil source, metallic particles from the refinery
equipment, and some dumped chemical wastes), plus some 0.5% water. The energy and material flows of partial oxidation of
heavy oil-based ammonia production are illustrated in Fig. 10. HFO can be cracked using gasification methods as illustrated in
Fig. 11 for hydrogen production and it can be coupled to ammonia synthesis plant.

3.2.2.1.3 Methane dissociation for ammonia production


In this case, methane is utilized as a source of hydrogen which is required for ammonia production as shown in Fig. 12. Rather
than SMR, thermal plasma cracking of ammonia is an important option for clean ammonia production. Methane (CH4) is
dissociated to carbon (C) black and hydrogen (H2) according to:
kJ
CH4 -C þ 2H2 ; DH ¼ 74:9 at 251C ð1Þ
mol
Methane is a favored option for hydrogen production from a hydrocarbon because of its high H to C ratio, availability and low-
cost. Furthermore, the C produced can be sold as a coproduct into the carbon black market which could be utilized in inks, paints,
tires, batteries, etc. or sequestered, stored, and used as a clean fuel for electricity production. The sequestering or storing of solid C
requires much less development than sequestering gaseous CO2.
As illustrated in Fig. 12, liquid natural gas and liquefied natural gas is pressurized using cryogenic pumps in this conceptual
system. The available excess heat from Haber–Bosch reactor is utilized for vaporization of both liquid nitrogen and liquefied
natural gas (LNG). The high-pressure LNG is sent to thermal plasma disassociation process which divides methane into hydrogen
and carbon. Here, different dissociation techniques such as microwave, non-thermal can also be used depending on the avail-
ability. Carbon black is stored for further utilization is various industries. The obtained hydrogen gas is transferred to Haber–Bosch
process together with high-pressure nitrogen gas. Here, any energy source can be used for electricity requirements, however
hydropower source is a clean and low-cost option especially for rich hydropower locations leading reduced total costs.

3.2.2.2 Water Electrolysis


In this method, the origin of hydrogen is water. Using electricity, the bonds of hydrogen can be decomposed and hydrogen can be
generated for ammonia production. There are many pathways for the electrolysis of water driven by wind, solar, geothermal, etc.
sources. Some of them are shortly explained here.

3.2.2.2.1 Wind electrolysis-based ammonia production


The system considered for producing hydrogen from wind energy involves two main devices: a wind turbine that produces
electricity, which in turn drives a water electrolysis unit that produces hydrogen. Wind energy is converted to mechanical work by
wind turbines and then transformed by an alternator to alternating current (AC) electricity, which is transmitted to the power grid.
The efficiency of wind turbines depends on location. Wind energy applications are normally preferred in areas with high wind
activity. Wind to ammonia systems shown in Fig. 13 produce ammonia through the use of electricity from wind turbine gen-
erators, which are usually large horizontal-axis wind turbines mounted on a tower. The energy and material flows in wind energy-
based ammonia production process are illustrated in Fig. 14. Wind turbines are commercially available in sizes up to about 3–8
MW of nameplate capacity for onshore applications and even larger machines can be found in offshore applications. The electrical
output of the wind turbine is highly dependent on wind speed, resulting in a high variability in electrical energy production. The
produced ammonia can serve as an energy storage medium and then can be converted into electricity appropriate power gen-
eration methods. The basic ammonia synthesis design is to use an electrolyzer to produce hydrogen from water and an air
separation unit to obtain nitrogen from air, both of which are combined in a Haber–Bosch synthesis reactor for production of
ammonia. Offshore wind-based ammonia is also an environmentally friendly solution for ammonia production, however, the
unit cost for ammonia is still higher.
50 Ammonia Production

Air
Heavy oil

O2
Cryogenic air
Gasification separation Electricity
Heat

Soot
removal/ Slag
recovery

Sulphur
removal/ Sulphur
recovery

Syngas

Shift
Water conversion Heat

Heat CO2 Condensate


Electricity removal CO2

H2
Flue gas N2
Liquid N2
wash
Heat and
flash gas

Electricity Ammonia Electricity


Compression synthesis

NH3

Ammonia
storage

NH3

Ammonia usage

Fig. 11 Ammonia production via partial oxidation of heavy oil.

3.2.2.2.2 Geothermal electrolysis-based ammonia production


Geothermal energy is a type of promising energy source for either heat driven applications or power generation systems by using
direct or indirect utilization processes. One of the advantages over other renewable energy sources is that geothermal energy is not
subject to high seasonal or daily deviations. A separation method of a geothermal fluid mixture is typically required for indirect
geothermal utilization, particularly in power generation systems. The separation process disposes of the liquid form of low
grade thermal energy which could be utilized further for other direct and indirect utilizations, such as a power plant bottoming
unit, heating and cooling purposes or other heat driven systems, depending on the remaining amount of energy. Geothermal
power plants can provide renewable and pollution-free electricity. The most active geothermal resources are usually found along
major plate boundaries where earthquakes and volcanoes are concentrated. Geothermal energy has the potential to play a
significant role in moving the world toward a cleaner, more sustainable energy generation and utilization options. Geothermal
power plants can be integrated into hydrogen production via water electrolysis which can be then converted to ammonia as shown
in Fig. 15.

3.2.2.2.3 Solar energy electrolysis-based ammonia production


Solar power is probably, along with wind power, the most readily available solution to clean energy alternatives. Solar cells
produce direct current electricity from light, which can be used to power the electrolyzers. Using the photovoltaic electrolysis
technology, hydrogen can be produced and reacted with nitrogen gas to produce ammonia as shown in Fig. 16. The energy and
Ammonia Production 51

Low cost hydropower

Hydroelectric
Electricity power plant

Excess heat

Electricity
Liquid Power High
natural pressure
gas (LNG) hydrogen
Haber−Bosch
Plasma Ammonia
Evaporator with no main compressor
dissociation (NH3)
ammonia synthesis

High
pressure High pressure
Excess
natural gas Carbon nitrogen
heat

Evaporator

Electricity

Liquid
nitrogen (N2)

Fig. 12 The schematic diagram of thermal plasma disassociation of methane integrated to ammonia production.

material flows in solar energy-based ammonia production are shown in Fig. 17. In addition to conventional Haber–Bosch process,
the electrochemical pathways can be easily integrated to renewable energy sources, particularly solar energy.

3.2.2.2.4 Nuclear-based ammonia production


Nuclear-based electricity yields lower cost and reliable supply. Combining nuclear power plant with ammonia production plant
is an encouraging method. In high-temperature electrolysis, the excess heat in the nuclear power plant is utilized to decrease
the required amount of electricity for electrolysis as seen in Fig. 18. Besides, in standard nuclear electrolysis-based option,
electricity is produced in nuclear power plant and directly utilized in electrolysis coupled with Haber–Bosch ammonia synthesis
loop. There is no heat assisting in this method. Hence, more electrical energy is required to split water into hydrogen and
oxygen. The schematic diagram of nuclear high-temperature electrolysis-based ammonia production option can be seen in
Figs. 18 and 19.
Steam electrolysis is the primary choice for hydrogen production in nuclear-based ammonia production. The efficiency
is comparable to the practical efficiencies of thermochemical processes if powered by an high-temperature gas-cooled
reactor (HTGR). Steam electrolysis can be powered by a pressurized water reactor (PWR) or a boiling water reactor (BWR).
HTGR-integrated high-temperature steam electrolysis (HTSE) can also be considered to supply hydrogen to the ammonia process.
The copper–chlorine (CuCl) cycle is a multiple step thermochemical cycle for the production of hydrogen. The CuCl cycle
is a combined process that employs both thermochemical and electrolysis steps. The CuCl cycle involves four chemical
reactions for water splitting, whose net reaction decomposes water into hydrogen and oxygen. Both heat and electricity are
provided at the same time for hydrogen generation and then hydrogen reacts with nitrogen to produce ammonia. A schematic
diagram of energy and material flows of nuclear thermochemical CuCl cycle-based ammonia production options are shown in
Fig. 20.
Nuclear-based ammonia production options can be cost competitive than other methods, however, the initial investment costs
are currently much higher. High capital costs are the major disadvantage of nuclear powered ammonia production however, by
enabling smaller and standardized modular reactors, it could reduce the capital costs, construction cost and time, and licensing
cost and time.
52 Ammonia Production

Air
Air

Wind turbine

Electricity
Electricity

Water Cryogenic air


Electrolyser O2 O2
separation

H2 Ammonia N2
synthesis

NH3

Ammonia
storage

NH3

Ammonia usage

Fig. 13 Ammonia production via wind energy-based electrolysis.

Wind (air)

Wind
Water turbine

Electricity

Nitrogen (N2) Wind electrolysis


Ammonia synthesis

Ammonia
(NH3)

Fig. 14 Energy and material flows in wind energy-based ammonia production.

3.2.2.3 Gasification
Coal and biomass are two of the materials which can be gasified for syngas production. After syngas is produced, hydrogen is
separated and used in ammonia synthesis plant. These two methods are shortly explained here.
Ammonia Production 53

Air
Geothermal

Geothermal to
electricity

Electricity
Electricity

Cryogenic air
Water Electrolysis O2 O2
separation

H2 Ammonia N2
synthesis

NH3

Ammonia
storage

NH3

Ammonia usage

Fig. 15 Ammonia production via geothermal energy-based electrolysis.

3.2.2.3.1 Coal gasification-based ammonia production


There are mainly two types of coal gasification. The one is called as underground coal gasification (UCG) which takes place below
earth level and the other one is coal gasification which takes place above earth level. Coal gasification is the second most
commonly used process for hydrogen production. With depletion of oil and gas resources the dependence on coal may increase
however the environmental issues are still challenging.
UCG is an encouraging option for the future use of unworked coal. Instead of mining coal reserves, UCG may ultimately make
unreached coal reserves accessible. UCG is one of the unmined type of electricity and other products generation by minimal GHG
emission. It prevents environmental effects, safety risks and health risks of mining. Carbon capture and storage of carbon dioxide
technology are treated as two effective technologies. Underground gasification is a process where coal, in place, is consumed by
partial combustion with air, oxygen, steam, or any combination of these to produce syngas. The syngas produced through the
gasification process consists mainly of hydrogen and carbon monoxide (CO). The obtained hydrogen is then used for ammonia
production. The simple layout of the system for ammonia production via UCG is shown in Figs. 21 and 22.

3.2.2.3.2 Biomass gasification-based ammonia production


The biomass gasification process provides a reliable and credible alternative and one of the fastest growing renewable technologies.
As an energy source, biomass can either be used directly via combustion to produce heat, or indirectly after converting it to various
forms of biofuel. Conversion of biomass to biofuel can be achieved by different methods which can be categorized into: thermal,
chemical, and biochemical methods. The energy and material flows in biomass DG-based ammonia production are illustrated in
Fig. 23. For biomass-based ammonia production plant, biomass is firstly pretreated and then sent to a downdraft gasifier and
pressure swing adsorption for hydrogen production as shown in Fig. 24. Generated hydrogen can be combined with nitrogen in
either Haber–Bosch process or novel electrochemical synthesis processes.

3.2.2.4 Novel Ammonia Production Methods


There are multiple pathways for ammonia synthesis besides mostly used Haber–Bosch process within the literature. For the
electrolytic routes, required hydrogen can be sourced from natural gas like the Haber–Bosch process or electrolysis of water, or
even decomposition of an organic liquid such as ethanol. When hydrogen is produced from water electrolysis utilizing a renewable
54 Ammonia Production

Solar Air

Photovoltaic

Electricity
Electricity

Cryogenic air
Water Electrolyser O2 O2
separation

H2 N2
Ammonia
synthesis

NH3

Ammonia
storage

NH3

Ammonia usage

Fig. 16 Ammonia production via solar energy-based electrolysis.

Solar irradiance

Solar
Water photovoltaics

Electricity

Nitrogen (N2)
Solar electrolysis
Ammonia synthesis

Ammonia
(NH3)

Fig. 17 Energy and material flows in solar energy-based ammonia production.

energy source, such as wind or solar, environmentally pollutant emissions would noticeably diminish for ammonia production
[15]. Water can also be utilized as a source of hydrogen inside the electrolytic cell through its reaction in the electrochemical
process. The use of water as a source of hydrogen would also be helpful in eliminating any issues of catalyst poisoning due
to traces of sulfur compounds or CO which are common impurities in hydrogen produced via steam reforming of natural gas.
Ammonia Production 55

Air
Uranium

Nuclear power
plant
Electricity
Electricity

Water High
Cryogenic air O2
Nuclear temperature O2
separation
heat electrolysis

H2 N2
Ammonia
synthesis

NH3

Ammonia
storage

NH3

Ammonia usage

Fig. 18 Ammonia production via nuclear high-temperature electrolysis.

Uranium

Nuclear power
Water plant

Heat Electricity

Nuclear high temperature


Nitrogen (N2) electrolysis
Ammonia synthesis

Ammonia
(NH3)

Fig. 19 Energy and material flows of nuclear high-temperature electrolysis-based ammonia production.

The process can be carried out under ambient conditions or at higher temperatures depending on the type of the electrolyte
material used. For high-temperature electrolytic routes of ammonia production, the use of waste heat from thermal or nuclear
power plants or heat from renewable energy sources like solar would make the overall process more environmentally friendly.
Ammonia production from hydrogen and nitrogen is exothermic in nature and is facilitated by high pressures and low tem-
peratures. Thus a balance between the operating temperature, pressure, and the ammonia yield needs to be proven for each
electrochemical system in determining ammonia production rates.
56 Ammonia Production

Uranium

Nuclear power
Water plant

Heat Electricity

Nuclear thermochemical
Nitrogen (N2) (CuCl)
Ammonia synthesis

Ammonia
(NH3)

Fig. 20 Energy and material flows of nuclear 3 step copper–chlorine (CuCl) cycle-based ammonia production.

Underground coal

Air

Underground coal
Air/O2 Water/steam
gasification

Syngas
expansion

Cryogenic air Electricity


Sulfur and Sulfur and separation
TAR removal TAR

Syngas to
Steam hydrogen
conversion

CO2
CO2 removal

Pressure
swing
adsorption

Electricity

H2 Ammonia N2
synthesis

NH3

Ammonia
storage

NH3

Ammonia usage

Fig. 21 Ammonia production via underground coal gasification (UCG) process.


Ammonia Production 57

Water/steam

Air/O2 Nitrogen (N2)


Underground coal gasification
Underground Ammonia synthesis
Electricity
coal

Ammonia
(NH3)

Fig. 22 Energy and material flows in underground coal gasification (UCG)-based ammonia production.

Steam Nitrogen (N2)


Biomass DG
Ammonia synthesis
Biomass Electricity

Ammonia
(NH3)

Fig. 23 Energy and material flows in biomass DG-based ammonia production.

Biomass
Air

Pretreatment

Steam Downdraft Cryogenic air


gasifier O2 separation Electricity

Pressure
swing
adsorption

Electricity

H2 Ammonia N2
synthesis

NH3

Ammonia
storage

NH3

Ammonia usage

Fig. 24 Ammonia production via biomass DG.


58 Ammonia Production

Electrolytic
ammonia synthesis

Liquid Composite Solid state


Molten salts
electrolyte membrane electrolyte

Organic solvents − Proton conducting


(Li, K, Cs) CI
LiCIO4 in membranes − Nafion
Eutectic with Li3N
tetrahydrofuran (RT − 80°C)
(180−500°C)
(room temperature)

− (Na, K, Li) carbonate Oxygen ion conducting


Ionic liquids −
and LiAIO2 ceramic membranes −
LiCIO4 in IL
(400−450°C) 8 mol% Y2O3−ZrO2
(room temperature)
− YDC-Ca3(PO4)2-K3PO4 (650°C)
(650°C)
Aqueous solutions Proton conducting
− Li2SO4 in 0.03M H2SO4 ceramic membranes
− LiCIO4 in 0.03M H2SO4 − Yb2O3 doped SrCeO3
(room temperature) − Sm2O3 doped BaCeO3
− Gd2O3 doped CeO2
(600−750°C)

Fig. 25 Main electrochemical ammonia synthesis electrolytes. Adapted from Giddey S, Badwal SPS, Kulkarni A. Review of electrochemical
ammonia production technologies and materials. Int J Hydrogen Energy 2013;38(34):14576–94.

There are four main categories of electrolytes used for electrochemical ammonia production as shown schematically in Fig. 25.
These are listed as follows:

• Liquid electrolytes which operate near room temperature.


• Molten salt electrolytes operating at intermediate temperatures (180–5001C).
• Composite electrolytes consisting of a traditional solid electrolyte mixed with a low melting salt (300–7001C).
• Solid electrolytes with a wide operating temperature range from near room temperature up to 700–8001C depending on the
type of electrolyte membrane used.

Electrochemical synthesis of ammonia is made possible by intensive research and advances in materials at the anode, cathode,
and electrolyte. The most significant developments have been the test with proton exchange membranes and specific combinations of
anode and cathode materials which have resulted in significant results. Xu et al. [16] investigated the synthesis of ammonia at
atmospheric pressure and low-temperature electrochemically, using the Sr2Fe1–xCoxNbO6–δ (SFCN) materials as the cathode, a
Nafion membrane as the electrolyte, samaria-doped ceria (SDC) as the anode and silver–platinum paste as the current collector.
Ammonia was produced from 25 to 1001C temperature levels when the SFCN materials were utilized as a cathode, with
SmFe0.7Cu0.1Ni0.2O3 which gives the maximum rates of ammonia formation. The highest rate of formation of ammonia was
1.13  108 mol/cm2/s1 at 801C, and the current efficiency is obtained as high as 90.4%. Other electrolyte-based systems have been
researched as well, and the results are shown in Ref. [17]. A comparison of current efficiency values from open literature is illustrated
in Fig. 26. Murakami et al. [18] proposed an electrolytic ammonia synthesis method from methane and nitrogen gasses in molten salt
under atmospheric pressure. Their experiments confirmed that ammonia was synthesized by potentiostatic electrolysis using a
methane gas electrode at 2.1 V and 773K. Another way of producing ammonia electrochemically is by reducing nitrate in aqueous
solutions. As discussed by Fanning [19], this nitrate is available from drinking water supplies and even nuclear waste. Not only is
ammonia produced in some cases but also hydrogen and nitrite with high reduction efficiencies with respect to current supply. A
comparative diagram showing the achieved current efficiencies of the electrochemical ammonia synthesis routes is illustrated in
Fig. 26.
Other considerations with the current state of the art are important to note. When hydrogen is used as a reactant in ammonia
synthesis, the source must be considered which is majorly of fossil fuels and a small portion from electrolysis which has an
efficiency of about 65–75%. Ammonia cracking is also required when supplying fuel cells that require pure hydrogen gas; this
must be done at temperatures above 5001C [20]. Li et al. [21] reported an appliance of electrochemical ammonia production
through an iron intermediate in which different electron transmission paths cogenerate H2 and NH3. At 200 mA/cm2, over 90%
of applied current drives hydrogen, rather than ammonia, formation. Lower temperature supports greater electrolyte hydration.
Ammonia Production 59

100
90
80

Coulombic efficiency (%)


70
60
50
40
30
20
10
0
Ceramic proton Liquid Molten salt Polymer
conductor electrolyte electrolyte membrane
based systems based systems based system based systems
Fig. 26 Achieved coulombic efficiencies of some electrochemical synthesis electrolytes. Data from Giddey S, Badwal SPS, Kulkarni A. Review of
electrochemical ammonia production technologies and materials. Int J Hydrogen Energy 2013;38(34):14576–94; Xu G, Liu R, Wang J. Electrochemical
synthesis of ammonia using a cell with a Nafion membrane and SmFe0.7Cu0.3  xNixO3 (x¼0  0.3) cathode at atmospheric pressure and lower
temperature. Sci Chin Ser B Chem 2009;52(8):1171–5 (SP Science in China Press); Murakami T, Nishikiori T, Nohira T, Ito Y. Electrolytic synthesis of
ammonia in molten salts under atmospheric pressure. J Am Chem Soc 2002;125(12):334–5; Licht S, Cui B, Wang B, et al. Ammonia synthesis by N2
and steam electrolysis in molten hydroxide suspensions of nanoscale Fe2O3. Science 2014;345(6197):637–40; Fanning JC. The chemical reduction of
nitrate in aqueous solution. Coord Chem Rev 2000;199(1):159–79; Li F-F, Licht S. Advances in understanding the mechanism and improved stability
of the synthesis of ammonia from air and water in hydroxide suspensions of nanoscale Fe2O3. Inorg Chem 2014;53(19):10042–4 (American Chemical
Society).

Table 1 Some novel studies for ammonia synthesis in the literature

Method Author Results

Molten salt electrolyte-based Licht et al. [23] Upwards of 30% current efficiency using water, and air as reactants. Using molten
system NaOH–KOH mixture with nanoscale Fe2O3 as a catalyst. Nickel monel mesh with
nickel electrodes
Liquid electrolyte-based systems Giddey et al. [17] Iron is used at the cathode at operating temperature of 501C when nitrogen was
supplied to cathode at a pressure of 50 atm producing 58% current efficiency
Ceramic proton conductor-based Giddey et al. [17] Ceramic proton conductor systems use vacancies in the chemical structures to
systems conduct charged species; the net conversion efficiency was found at about only
50% because of the decomposition of ammonia back to hydrogen and nitrogen
Polymer membrane-based Garagounis et al. [24] The highest product yield was demonstrated from a solid Nafion membrane with a
systems mixed oxide (SmFe0.7Cu0.1Ni0.2O3). The rate of ammonia formation reported was
1.13  10  8 mol/cm2s, obtained at 801C

To synthesize ammonia by electrolysis in hydroxide, water, N2 (or air), and nanoscopic Fe2O3 are simultaneously required. Lan
et al. [22] reported an artificial ammonia synthesis bypassing N2 separation and H2 production phases. A maximum ammonia
production rate of 1.14  105 mol/m2s was realized when a voltage of 1.6 V was applied. They implied that in the future, other
low-cost ammonia synthesis catalysts, such as Co3Mo3N and Ni2Mo3N41, could be used to exchange Pt for selective ammonia
synthesis. A brief description of the novel studies for ammonia production is given in Table 1.
A variety of factors needs to be analyzed when selecting the cell material, such as system operating temperature, current density,
pressure, and conductivity which all affect the ammonia production rate. It is important to note that conductivity of a solid
electrolyte increases exponentially with temperature and by reducing cell thickness as reported by Giddey et al. [17].
Garagounis et al. [24] summarized the test results of studies within the last 15 years using electrolyte cells. More than 30
electrolyte materials with 15 catalysts that were used as working electrodes (cathode) were tested. The polymer Nafion yielded
the highest rate of ammonia formation at a very low-temperature. Nafion can also be used as a proton conductor with and a
Ru/C cathode which yielded NH3 from H2O and N2 at 901C. These low operating temperatures are sought after when designing
new systems because of the reduced energy input required and a lower rate of decomposition of the ammonia formed. As an
alternative approach, the use of oxygen-ion (O2) conductors where steam and nitrogen are introduced together at the cathode
should be considered. The rate of NH3 production was however low in a demonstration at 5001C but improved by up to two
orders of magnitude at higher temperatures as reported by Skodra et al. [25]. Furthermore, NH3 synthesis using molten salt
electrolyte-based systems can yield high conversion ratios similar to that of polymer-based membranes.
60 Ammonia Production

Serizawa et al. [26] reported conversion ratios as high as 70% of Li3N into NH3 using a molten LiCl–KClCsCl system at
temperatures between 360 and 3901C. These conversion ratios have been achieved despite the side reactions where parts of NH3
were dissolved in the melt in the form of imide (NH2) and amide (H2N) anions resulting in a lower NH3 yield.

3.2.2.4.1 Liquid electrolyte-based systems


In this approach, Lithium perchlorate (LiClO4 (0.2 M)) in tetrahydrofuran as the electrolyte and ethanol (0.18 M) as the hydrogen
source can be used. In the preceding studies [27], a very low current efficiency of 3–5% was realized considering that the current
density was also low (2 mA/cm2). The current efficiency may improve under a different set of experimental circumstances by
varying the pressure and temperature values, but, under the conditions of the test, the breakdown of the ionic liquid electrolyte
was observed representing plain distresses about the long-term capability of the procedure. Also, the solubility of Li salts has been
stated to be low in many ionic fluids [27].
The synthesis of ammonia from electrochemistry is based on electrolysis where electric current is supplied to a reactor
containing a cathode, anode, and ionic conducting membrane. The chemical reactions contain reduction on one side and
oxidation on the additional with an essential constituent being the membrane/electrolyte which will only conduct a special kind
of ion such as protons in the form of H þ ; this lets the device to work uninterruptedly. There are many different
kinds of cathodes, anodes, and membranes/electrolytes though the principal remains the same where two reactants and an
activation potential are applied to produce a chemical reaction resulting in ammonia synthesis. These reactions have been shown
to work in an extensive variety of pressures and temperatures making it feasible when employed with atmospheric circumstances.
The present studies in molten salt-based electrochemical procedures have made some innovative developments. Using water
and atmospheric air, uniting them into a molten salt of NaOH–KOH with nano-Fe2O3 as the catalyst to achieve a 30% current
efficiency was experientially succeeded [23]. The systems were using an open pot system with no separator, two nickel conductors,
and a nickel monel mesh. Though the efficiency appears to be inferior to the proton conducting membranes, its supplies are cheap
and could see enhancements if the device set up is improved. In this way, the conductors and mesh required continuous inspection
and substituted owing to corrosion in the device. The electrolyte salts must be tested periodically for reactivity. When including a
catalyst inside the mix such as nano-iron oxide [23], the particles must be reserved in a strong concentration throughout the salt
mix. In order to realize this technique, a number of ideas have been established on minor cells in laboratory experiments using
hydrogen, methane or water as the reactants [18].
The technology is at the stage of development with a good current efficiency of 72% reported for hydrogen oxidation reaction
and ammonia synthesis rates.
The reactor shown in Fig. 27 has water and air/nitrogen being supplied, water being the source for the hydrogen in the
ammonia and air with the nitrogen. The molten salt begins to melt at around 1701C and at 2001C, it is in a molten form. High
temperatures in the range of 200  3001C must be sustained, and a corrosive resistant material should be used.

3.2.2.4.2 Composite membrane-based systems


The complex electrolytes entail of one or additional different ionic conducting stages and the second or third phase is added to the
main stage to adjust electrical, thermal, or mechanical properties. For instance, an alkali metal carbonate and an oxide, such as
LiAlO2 or Sm2O3 doped CeO2, have been shown to have oxygen-ion, carbonate ion, and even proton conductivity under
certain situations (e.g., in the attendance of hydrogen) [28,29]. Such materials have been under investigation as possible

+ –
Electricity

H2O N2

H2
Unreacted
N2
Waste heat

NH3+H2
Condenser Molten salt reactor O2
~200°C

NH3

Heat
Fig. 27 Steam-based ammonia synthesis in molten salt reactor.
Ammonia Production 61

electrolytes for intermediate temperature (400–8001C) fuel cells and are also being employed to study ammonia manufacture
rates under a range of effective conditions.

3.2.2.4.3 Solid state electrolyte


A number of different systems, based either on proton or mixed proton/oxygen-ion-conducting solid electrolytes, are experiencing
investigation and growth for application in electrochemical ammonia synthesis. The main apparatuses of the solid-state elec-
trochemical device are two porous electrodes anode and cathode detached by a dense solid electrolyte, which lets ion conveyance
of either protons or oxide ions and supports as a fence to gas diffusion [30,31]. Solid-state proton conductors (SSPC) denote a
class of ionic solid electrolytes which have the competence to transmission hydrogen ions (H þ ) [32]. Though, this method has
some difficulties, such as high-temperature materials and creation of secondary phases [33,34]. A schematic drawing of SSAS is
shown in Fig. 28.
The SSAS system can be coupled to photoelectrochemical hydrogen production as illustrated in the Fig. 29. In the electrolytic
cell, two metal electrodes are placed on both sides of the proton conductor. The gaseous H2 passing over the anode will be
converted to H þ . Hydrogen in the form of protons will be transported to the cathode where the half-cell reaction occurs.

3.2.2.4.4 Ceramic/inorganic proton conducting solid electrolyte-based systems


A characteristic electrolytic cell for ammonia synthesis is obtained by depositing electrode (catalyst) coverings on both sides of the
proton conducting membrane. These absorbent conductors are characteristically screen printed or brush coated on ceramic proton
electrode membranes followed by heat treatment. Water or hydrogen is fed to the anode and nitrogen to the cathode, and
ammonia is produced on the cathode side of the cell. The current collection is attained by insertion metallic meshes or pieces in
interaction with these conductors. Obviously, the proton conducting ceramic membrane, along with cathode or ammonia
synthesis catalyst, are most significant mechanisms in these schemes. These membranes are obligatory to reveal considerable
proton conductivity at temperatures above 4001C [17].

3.2.2.4.5 Polymer membrane-based systems


There are numerous kinds of polymer ion exchange membranes available that can be used as an electrolyte in electrochemical
ammonia synthesis cells. These membranes can be functioned in the temperature variety from room temperature to 1201C. Nafion
membranes are the greatest popular proton conducting membranes used in the chlor-alkali industry, and in the polymer elec-
trolyte membrane (PEM)-based fuel cells and electrolysis cells [17]. Notwithstanding some stability issues for polymer membranes
in the attendance of ammonia, there are numerous compensations of using these membranes due to their high proton con-
ductivities at inferior temperatures and a large quantity of information available for cell building and assembly owing to their

High temperature ~ 600˚C


Electricity
Unreacted N2

C
H+ a
A t
n h
Unreacted

o
Electrolyte

o
H2O

d d NH3
e e

O2

H2O (g) N2

Fig. 28 Steam-based solid state ammonia synthesis (SSAS) in a high-temperature electrochemical ammonia synthesis cell.
62 Ammonia Production

Solid state ammonia synthesis


(H2 based) cascaded to PEC
Electricity

H+ C
A
a
n

Electrolyte
t
Unreacted o NH3
h
H2 d
o
e
d
e
Solar
Unreacted
irradiation Electricity
N2

PEC reactor
Water H2 N2

O2

Fig. 29 Solid state ammonia synthesis (SSAS) using photo-electrochemically generated hydrogen.

practice in fuel cells. The low-temperature process would decrease the degree of decomposition of ammonia formed and evade
numerous extra high-temperature resources related issues.

3.2.2.4.6 O2 conducting membrane materials and ammonia synthesis systems


A big number of O2 conducting ceramic electrolytes is available and have been used in oxygen sensors, solid oxide fuel cells and
for HTSE. These contain completely and partly steadied ZrO2 with different dopant kinds and levels, doped CeO2, doped LaGaO3
at A- and B-sites, and doped Bi2O3 [35]. The O2 conductivity feature differs meaningfully with the kind of material used.
Although some of these materials retain O2 conductivity over an extensive variety of temperatures, oxygen partial pressures, and
gas compositions, others grow either electronic or even proton conductivity in the attendance of water and similar to those stated
for doped BaSrO3 and SrCeO3 [33].
Characteristically, electrochemical ways, examined so far need a process at abundant lower pressures than those used in the
Haber–Bosch procedure with working temperatures from close room temperature for liquid and polymer electrolyte systems to
between 400 and 8001C for other solid electrolytic ways. The low-temperature process has the potential to decrease material and
working costs and upsurge life period of the electrochemical device providing high ammonia manufacture rates and high coulombic
efficiency can be attained. One significant benefit of ammonia production by high/moderate-temperature electrolytic ways is that
such systems can be combined with renewable energy, thermal or nuclear power plants to deliver the waste heat for high-temperature
process thus reducing the overall energy input particularly if water is used as the hydrogen source [22,35,36]. There is also solar
thermochemical ammonia synthesis systems. Latest studies on solar thermochemical ammonia manufacture also display that a net
efficiency ranging from 26 to 33% can be obtained by combining the ammonia synthesis cycle with hydrogen generation [37].

3.2.3 Case Study 1: Electrochemical Ammonia Production

The electrochemical process can be carried out under ambient conditions or at higher temperatures depending on the type of the
electrolyte material used. Here, a case study is performed for electrochemical synthesis of ammonnia using hydrogen and nitrogen
as the reactants.
Although synthesis of NH3 using water as a hydrogen source in an electrochemical process is attractive and reduces additional
step, for the cases where ammonia and hydrogen are individually required as alternative fuels, H2 can be directly utilized in the
electrochemical NH3 formation as investigated in this study. Most of the literature used water as hydrogen source which also
requires water splitting process at the same time with ammonia synthesis. Specifically, for solar energy storage applications, H2 can
act as short-term storage whereas NH3 can serve as a long-term storage medium which reduces the storage losses significantly. In
this case study, we report the electrochemical synthesis of ammonia using H2 and N2 at ambient pressure in a molten hydroxide
ambient with the nano-Fe3O4 catalyst. The active surface areas of the Nickel mesh electrodes are increased to allow higher
formation rates. The effects of various parameters, such as applied potential and current density, reaction temperature on ammonia
formation rate are investigated. The nickel mesh electrodes are utilized having a large surface area of 100 cm2 and the reaction
temperatures are quite lower than the conventional Haber–Bosch process.
Ammonia Production 63

3.2.3.1 Electrochemical Ammonia Synthesis


In this case study, H2 and N2 are directly used for electrochemical synthesis of ammonia at the electrodes. N2 receives the electrons
from external power supply. Hence nitrogen gas sent via the porous nickel cathode is reduced to nitride according to the following
equation:
N2 þ 6e -2N3 ð2Þ
3
It becomes N then after moves to the other electrode where H2 is being supplied. Hydrogen ions combine with nitrogen ions
and form NH3 at anode electrode as illustrated in Fig. 30 and shown the following equation:
2N3 þ 3H2 -2NH3 þ 6e ð3Þ
The anode reaction is also achieved on porous nickel electrode.
The overall reaction is:
3H2 þ N2 -2NH3 ð4Þ
Hydrogen and nitrogen are required separately to be produced and supplied to the ammonia synthesis reactor. To conduct the
electrochemical reaction for ammonia synthesis, reactant nitrogen is supplied from the nitrogen tank. For the production of
hydrogen, a separate electrode–electrolyte assembly is formed consisting of graphite rods and NaOH electrolyte. The volume of the
electrolyte is 1 L whereas the molarity of NaOH solution is 1 M.
Nickel mesh is used for both electrodes each having an area of 100 cm2 as shown in Fig. 31. Nickel meshes have high melting
point, non-corrosivity, high conductivity and excellent stability in the molten salt medium. The area of 100 cm2 is used for
Faradaic efficiency calculations. The reactor, 500 mL crucible, is made of Alumina (Al2O3) being 99.6% pure, having high melting

e– e–

NH3

e–
e–
e– H
Ni electrode

H2 N3– N3– N2
Ni electrode

Molten salt
electrolyte

Fig. 30 Electrochemical ammonia synthesis reaction in molten salt medium.

Fig. 31 Nickel mesh electrodes in the reactor for the reactor.


64 Ammonia Production

Fig. 32 Heating tape used around the alumina crucible.

point, strong hardness, chemical stability, and non-corrosivity. The cover plates are made of stainless steel (316 alloys) which
withstand high temperatures.
The molten salt electrolyte is a mixture of 0.5 M NaOH and 0.5 M KOH. The mass of the NaOH is 221 g whereas KOH mass is
310 g. The total volume of the mixture was about 430 mL at 2001C. The mixture is originally prepared at room temperature,
putting the salts into the reactor to melt in the crucible when initially heated up to 2551C. The experiments were performed at 210
and 2151C using the heating tape shown in Fig. 32.
Iron oxide (Fe3O4) as nano-powder (20–30 nm, 98 þ %) is used in the experiments as a catalyst. The high surface area of the
nano-Fe3O4 in the electrochemical synthesis is critical for the reaction to occur and to obtain higher ammonia evolution rates.
Since ammonia is highly soluble in water, the molten salt electrolyte is not mixed with the water inside the reactor to allow higher
ammonia capturing in the H2SO4 solution.
The ammonia electro-synthesis chamber comprises a nickel mesh cathode and a nickel mesh anode immersed in molten
hydroxide electrolyte containing 10 g suspension of the nano-Fe3O4 contained in alumina crucible sealed to allow gas inlet
at the cathode and gas outlet from the exit tubes. The reactants, H2 and N2, are bubbled through the mesh over the anode
and cathode, respectively. The combined gas products (H2, N2, and NH3) exit through two exit tubes in chamber head space.
The exiting gasses are firstly measured using flow meters and bubbled through an ammonia water trap then analyzed for
ammonia, and subsequently, the NH3 scrubbed-gas is analyzed for H2 or N2 using hydrogen analyzer device (ABB Continuous
gas analyzers model AO2020). In the alumina crucible cell, the anode consists of a pure Ni mesh with an area of 100 cm2 and the
100 cm2 cathode composed of the same material. These Ni meshes are stable in the molten 200–2501C hydroxide. The electrodes
are connected externally by spot welded Ni wires. The reactor is kept at constant temperature using on/off type temperature
controller, and the internal temperature of the reactor is continuously measured using a Pt 100 temperature probe inside the
reactor body. As mentioned earlier, the product gasses from the reactor is bubbled through an ammonia trap consisting of a dilute
500 mL 0.001 M H2SO4 solution, changed every 15 min for ammonia analysis. Ammonia concentration is determined using
various techniques to confirm the results. The methods utilized are as follows: ammonia test strips, ammonia gas flowmeters,
Arduino ammonia gas sensor, and salicylate-based ammonia determination method. For the salicylate-based method, two dif-
ferent solutions are used where one of them contains sodium salicylate, and the other one contains sodium hydroxide and sodium
hypochlorite. In each case, redundant measurements yield similar ammonia formation values, with the observed reproducibility of
methodologies. Also, the pH level of the dilute H2SO4 solutions is recorded before and after NH3 trapped in the solution in order
to observe the dissolved ammonia. Ammonia formation rate is calculated by converting the measured NH3 to moles per seconds
and considering the surface area of Ni electrodes as 100 cm2. The ammonia formation rate is calculated using the following
equation:
 
NHþ4 V
m_ NH3 ¼ ð5Þ
t

where [NHþ 4 ] is the concentration of formed ammonia in mg/L, V is the total volume of H2SO4 for trapping ammonia as L and t is
the time of collection.
Ammonia Production 65

The Faradaic efficiency is calculated based on the moles of electrons consumed compared to the 3e/NH3 equivalents pro-
duced. Thus, the Faradaic efficiency of ammonia generation process is defined as follows:
n_ NH3  F  3
ZFaradaic ð%Þ ¼ ð6Þ
i
where F is Faraday constant and i is the current density (A/cm2).
The energy efficiency of the ammonia production process is also calculated based on lower heating values (LHVs) of reacted
hydrogen and ammonia, nitrogen enthalpy and electrical power input as follows:
_ NH3  LHV NH3
m
ZEnergy ð%Þ ¼ ð7Þ
_ H2  LHV H2 þ m
ðm _ el Þ
_ N2  hN2 þ W
_ el is the total electricity input during the experiment calculated using the total charge, applied voltage and duration.
where W

3.2.3.2 Case Study 1 Results and Discussion


The pure alkali hydroxides NaOH and KOH each melt only at temperatures above 3001C. The individual melting temperatures
of NaOH and KOH are 318 and 4061C, respectively. Among various alternatives, these two salts melt at quite lower temperatures
which are a highly desired property in order to decrease external heat energy input. Based on common materials, the NaOH–KOH
eutectic is of particular attention and melts at 1701C. Ammonia synthesis rates increase when the molten hydroxide (NaOH–
KOH) electrolyte is mixed with high surface area Fe3O4 to provide iron as a reactive surface and when nitrogen and hydrogen
are present in the reactor. The molten salt medium is supplied electricity between two nickel anode and cathode electrodes. The
mixture is prepared in the beginning by simply adding NaOH and KOH pellets in the reactor. After the salts melt, nano-Fe3O4 is
added to the electrolyte and then stirred. When the mixture is ready, the lid is tightly closed and sealed. In order to yield NH3 in the
reactor, H2, N2 and nano-Fe3O4 are simultaneously needed.
Table 2 shows the experimental conditions. The experiments are performed at constant current in galvanostatic mode. The
temperatures given in the table are average temperatures because the temperature controller is on/off type and keeping the
temperature constant causes fluctuations. For each run, different ammonia trapping H2SO4 solution is used. The unreacted H2 is
also measured using a hydrogen sensor embedded to Arduino board which shows the portion of H2 which does not react. Table 3
shows the summary of the results showing the NH3 production rates and calculated efficiencies.
The required cell voltage to initiate the reaction of nitrogen and hydrogen in molten hydroxide at 2101C in the existence of
nano-Fe3O4 is measured to be on average 1.4 V when the applied current is 200 mA between the 100 cm2 Ni electrodes in the
molten NaOH–KOH electrolyte as depicted in Fig. 33. At 2 mA/cm2 and 2101C, ammonia is synthesized at a rate of 6.54  1010
mol/s cm2. At 2151C in a eutectic Na0.5K0.5OH electrolyte with suspended nano-Fe3O4, at 2 mA/cm2 applied current, NH3 is
generated at a Faradaic efficiency of about 6.3%.
The NH3 synthesis rates are obtained within the first hour of the experiments as seen in Fig. 34. Further studies to enhance the
stability of NH3 generation rate are important to note. It is also observed in the experiments that when the reactor temperature is
below 2001C, ammonia is generated with a similar production rate to above 2001C. The fluctuations in the potentials are caused
by the temperature on/off process to keep the temperature constant during the experiments. It is observed in the experiments that
lower current density and lower temperature improve the stability of the rate of NH3 evolution.
The measured Faradaic and energetic efficiencies of ammonia evolution in time at different temperature levels and conditions
in NaOH–KOH molten electrolyte are comparatively illustrated in Fig. 35. The generated NH3 is trapped and measured in a room
temperature dilute H2SO4 trap. A non-dilute H2SO4 trap is also tried before the experiments reported here to understand the
absorptivity of the solution. However, the ammonia readings were not successful in this case. Hence, dilute H2SO4 solutions are
utilized for the reported experiments. The conversion efficiency is not only dependent on the hydrogen amount but also the

Table 2 Experimental conditions

Experiment # Duration (min) Temperature (1C) Voltage (V) Current density (mA/cm2)

1 15 210 1.4 2
2 45 215 1.3 2

Table 3 Summary of the experimental results showing the NH3 formation rates and efficiencies

Experiment # NH3 formation NH3 mass flow Faradaic Energy


rate (mol/s cm2) rate (g/min) efficiency (%) efficiency (%)

1 6.54  10  10 6.67  10  5 9.45 6.66


2 1.91  10  10 1.94  10  5 2.76 2.07
66 Ammonia Production

1.5

1.4

1.3
Voltage (V)

1.2
Run #1 (2 mA/cm2 and 210°C)
Run #2 (2 mA/cm2 and 215°C)
1.1

1.0
0 200 400 600 800
Time (s)
Fig. 33 The relationship between voltage and time during two experimental runs at different applied currents and temperatures for
electrochemical synthesis of NH3 using N2 and H2 with nano-Fe3O4 in a molten salt hydroxide electrolyte.

0.14 2.5
Cumulative ammonia production (mmol)

Cumulative ammonia production (mg)


0.12
2.0
0.10

1.5
0.08
mol
0.06 mass
1.0

0.04
0.5
0.02

0.00 0.0
10 20 30 40 50 60
Time (min)
Fig. 34 Cumulative NH3 production amount by electrochemical synthesis using N2 and H2 with nano-Fe3O4 in a molten salt hydroxide electrolyte.

amount of catalyst available to stimulate the conversion of N2 and H2 into NH3. In order to make sure that there is enough N2 to
be reacted with supplied H2, the supplied volume of N2 is kept quite higher than H2.
The variations of NH3 formation rates at 2 mA/cm2 current density, and different temperature levels are comparatively shown in
Fig. 36. The figure reveals that the temperature and current density are not the sole parameters affecting the performance of the reaction.
At 2 mA/cm2 and 2101C, the NH3 formation rate is high corresponding to about 6.6  1010 mol/s cm2. The differentiations
might be caused by the catalyst saturations as well as the changes in supplied H2 rates.

3.2.3.3 Case Study 1 Conclusions


The electrochemical synthesis of NH3 is a promising alternative to conventional energy intensive NH3 production plants. Using
renewable energy resources to drive the electrochemical NH3 synthesis, the carbon footprint of current NH3 production industry
can be lowered significantly. Electrochemical NH3 synthesis routes offer higher integrability to stand alone and distributed NH3
production which is a carbon-free fuel for various sectors. In this case study, NH3 is electrochemically generated at ambient
Ammonia Production 67

10
Run 1
Run 1 Run 2
9

7 Run 1

Efficiency (%) 6

4
Run 2
3
Run 2
2

0
Faradaic efficiency (%) Energy efficiency (%)
Fig. 35 Faradaic and energy efficiencies of two experimental runs for electrochemical NH3 synthesis using N2 and H2 with nano-Fe3O4 in a
molten salt hydroxide electrolyte.

7 216

215
6
cm2)

214

Reaction temperature (°C)


(10−10 mol/s

5
213

4 212
mol/s cm2 Temperature
NH3 production

3 211

210
2
209
1
208

0 207
2 2

Current density (mA/cm2)


Fig. 36 Change of electrochemical NH3 formation rates depending on the applied current densities and reactor temperature using N2 and H2 with
nano-Fe3O4 in a molten salt hydroxide electrolyte.

pressure without a necessity of huge compressors using H2 and N2 in a molten hydroxide medium with the nano-Fe3O4 catalyst.
The reaction temperature is varied to investigate the impact of temperature on NH3 production rates. Having noncorrosive and
high surface area nickel mesh electrodes allowed to generate more NH3. The maximum Faradaic efficiency is calculated as 9.3% at
the reaction temperature of 2101C. The NH3 formation rate is determined to be 6.53  1010 mol/s cm2 at 2 mA/cm2 current
density. The possible issues being faced in the molten salt electrolyte-based electrochemical NH3 synthesis is expected to be further
resolved by way of not only the addition of more appropriate additives but also the continuous optimization of reactor
configuration.

3.2.4 Case Study 2: Life Cycle Assessment of Various Fuels Including Ammonia

Life cycle assessment (LCA) presents a systematic set of procedures for collecting and examining the contributions and productions
of materials and energy, and the related ecological impacts, directly assignable to a product, system or service throughout the
course of its life cycle. LCA is mainly a cradle to grave examination technique to inspect ecological influences of a scheme or
68 Ammonia Production

Goal and scope definition

Interpretation
Inventory analysis

Impact assessment

Improvement analysis

Fig. 37 Framework of life cycle assessment (LCA) analysis. Adapted From ISO 14040. Environmental management – Life cycle assessment –
Principles and framework. Available From: http://www.iso.org/iso/catalogue_detail?csnumber=37456; 2006 [accessed 07.01.17].

procedure or product based on the exact steps illustrated in Fig. 37. A life cycle is the set of stages of a product or service system,
from the extraction of natural resources to last removal. The general ecological impact of any process is not complete if the only
operation is considered, all the lifetime steps from reserve extraction to removal throughout the lifetime of a product or process
should be considered. Mass and energy flows and ecological impacts connected to plant building, use, and dismantling phases are
taken into account in LCA analysis [38].

3.2.4.1 Life Cycle Assessment Steps


The goal definition and scoping describe the product, process or action. It classifies the limits and environmental effects to be
deliberated for the evaluation. Inventory section categorizes and counts energy, water, and materials usage and ecological dis-
charges. Impact assessment stage assesses the human and environmental effects of energy, water, and material usage and
the ecological releases identified in the inventory analysis. Interpretation stage evaluates the consequences of the inventory
analysis and impact evaluation to choose the selected product, procedure or service. The events for performing of an LCA have
been defined by such International Organization for Standardization (ISO) based on ISO 14040 – Environmental management –
LCA – Principles and framework [39] and ISO 14044:2006 – Environmental management – LCA – Requirements and guidelines
[38]. The four steps are explained as follows:

3.2.4.1.1 Goal and scope definition


This is the primary stage of an LCA study. This stage describes the objectives of the study and also the range of activities under
examination. The highest care and feature is required to outline the goals and scope of the study. The LCA is an iterative procedure,
so the feedback consideration must be reserved in the definition of systems. Usually, sketch boundary and the energy designate the
scope and resources are considered for the processes falls within these boundary limits.

3.2.4.1.2 Inventory analysis


In this stage, raw material, and energy, the releases and waste information is collected. This data is used to compute the total
releases from the system. The mass and energy balances are used at each stage to compute the life cycle inventory (LCI) of the
system. The LCI needs to comprise each possible energy and material input and all possible releases to establish trustworthy
results. Data excellence is a significant aspect of LCA. Throughout inventory examination, the standards are followed for main-
taining the data excellence.

3.2.4.1.3 Impact assessment


The third step of LCA is life cycle impact assessment. This phase measures the influences of activities under investigation. The
LCI information is utilized to find out the affected zones. The LCI information is fundamentally analyzed in a two-step
procedure:
Classification: The Impact groups are established, and LCI information is analyzed to mark the information and compute the
values of releases corresponding to each group. The impact categories are based on the assessment technique used. As an instance,
some of the groups are global warming potential (GWP), acidification, human toxicity, etc.
Characterization: It is the second stage of assessment. Classification stage group the information in respective impact
groups. Characterization stage is used to evaluate the comparative contribution of each kind of emission to these impact
categories.
Ammonia Production 69

Normalization and weighting: This stage is not obligatory according to the standards. The releases are normalized
conforming to a typical and converted into a scoring system. The total score is used to classify the methods and processes of
concern.

3.2.4.1.4 Interpretation of results and improvement


The last stage is interpretation of results and feedback for improvement of the system. The unclear parts of the system are
recognized, and extremely contaminating procedures can be eliminated with cleaner replacements.
LCA can classify critical points where process changes might meaningfully decrease impacts.
The accomplishment of an LCA brings some advantages as follows:

• Assessing systematically the environmental consequences connected with a given product or process.
• Assessing the human and ecological effects of material and energy consumption and ecological emissions to the local com-
munity, district, and biosphere.
• Classifying impacts connected to specific environmental areas of concern.
• Assisting in classifying noteworthy changes in environmental impacts between life cycle phases and ecological media.
• Comparing the health and environmental impacts of substitute products and procedures.
• Quantifying ecological releases to air, water, and land in relation to each life cycle phase and main contributing process.

There is a number of assessment methods progressed over the time to classify and describe the environmental impacts of the
system. LCA can be applied by means of CML (Center of Environmental Science of Leiden University) 2001 technique which was
proposed by a set of scientists under the principal of CML including a group of impact classes and characterization procedures for
the impact assessment stage. Some of the baseline indicators of CML technique which are used in this chapter are explained as
follows [40]:

3.2.4.1.4.1 Depletion of abiotic resources


The key concern of this category is the human and ecosystem health that is affected by the extraction of minerals and fossil as
inputs to the system. For each extraction of minerals and fossil fuels, the abiotic depletion factor (ADF) is defined. This indicator
has globe scale where it is related to concentration reserves and rate of de-accumulation.

3.2.4.1.4.2 Human toxicity


Toxic substances on the human environment are the core concerns for this category. In the working environment, the health risks
are not included in this category. Characterization factors, human toxicity potentials (HTP), are determined with the uniform
system for the evaluation of substances (USES)-LCA, describing fate, exposure and effects of toxic substances for an infinite time
horizon. 1,4-Dichlorobenzene equivalents/kg emissions are used to express each toxic substance. Depending on the substance, the
geographical scale differs between local and global indicator.

3.2.4.1.4.3 Ozone depletion


Due to stratospheric ozone depletion, a bigger portion of UV-B radiation spreads the world surface. It may have damaging
properties upon human health, animal health, terrestrial and aquatic ecosystems, biochemical cycles and on materials. The
category is output related, and it is at a global scale. The model of characterization is advanced by the World Meteorological
Organization (WMO) and describes ozone depletion potential of various gases in a unit of kilogram CFC-11 equivalent/kg
emission. The geographic scope of this indicator is at a global scale and the span of time is infinity.

3.2.4.1.4.4 Global warming


The GHGs to air are associated with the climate change. Adversative effects upon ecosystem health, human health, and material
welfare can result from climate change. The intergovernmental panel on climate change (IPCC) developed the characterization
model which is selected for the development of characterization factors. A kilogram carbon dioxide/kilogram emission is used to
express the GWP for time horizon 500 years (GWP500). This indicator has a global scale.
The Eco-indicator method specifies the environmental impact in terms of numbers or scores. It simplifies the interpretation of
LCA by including a weighting method. After weighting, it supports to give a single score for each of the product or processes which
is calculated based on the relative environmental impact. The score is represented on a point scale (Pt), where a point (Pt) means
the annual environmental load (i.e., whole production/consumption undertakings in the economy) of an average citizen. Eco-
Indicator 99 (E) uses a load of average European [40]. The Eco-indicator 99 defines the environment damage in three broad
categories:

3.2.4.1.4.5 Human health


It includes the number and duration of diseases and loss of life years due to permanent deaths caused by environmental
degradation. The effects are included mainly by climate change, ozone layer depletion, carcinogenic effects, respiratory effects, and
ionization.
70 Ammonia Production

3.2.4.1.4.6 Ecosystem quality


This category includes the impact of species diversity, acidification, ecotoxicity, eutrophication, and land-use.

3.2.4.1.4.7 Resources
This category corresponds to the depletion of raw materials and energy resources. It is measured in terms of the surplus energy
required in future for the extraction of lower quality of energy and minerals. The agricultural resource depletion is studied under
the category of land use.

3.2.4.2 Case Study 2 Results and Discussion


SimaPro 7 software is used for LCA analyses. The two methods utilized for the current LCA analyses are CML 2001 and Eco-indicator
99. Human toxicity can play a major role for decision of using alternative methods of ammonia production. Besides, GWP is the key
features to compare the total CO2 equivalent emission from any source. Abiotic resources are natural resources including energy
resources. Since fossil fuels resources are declining gradually, abiotic depletion potential is a significant category for LCA analysis.
Fig. 38 shows the comparison of ozone layer depletion values for various transportation fuels. Ammonia has lowest ozone
layer depletion even if it is produced from SMR and partial oxidation of heavy oil. In addition, production of fuel ammonia yields
lower GHG emissions compared to petrol and propane production as shown in Fig. 39.

5.E-07
4.E-07
(kg CFC-11 eq/kg fuel)

4.E-07
Ozone layer depletion

3.E-07
3.E-07
2.E-07
2.E-07
1.E-07
5.E-08
0.E+00
y

nt

nt
an

an
er

er

er

la

la
in

in

fin

pl

pl

tp

tp
ef

ef

re

at

at

,a

,a
tr

tr

d,

d,
at

ng

gy
,a

,a

ui

ui
r,

er
ki
ne

ed

liq

liq
fu

ac

en
ta

ad

ul

g,

n,

cr
bu

-s

d
le

in

tio

in
w

on
un
e/

rm

da
lo

,w
an

rb
fo
l,

xi
l,

ia
ca
tro
op

se

re

lo

on
ro
Pe

ie
Pr

am

tia

m
yd
D

ar

Am
te

,h
,p
,s

ia
ia
ia

on
on
on

m
m
m

Am
Am
Am

Fig. 38 Ozone layer depletion during productions of various fuels.

Diesel, low-sulfur, at refinery

Ammonia, from wind energy, at plant

Propane/butane, at refinery

Petrol, unleaded, at refinery

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7


Global warming potential (kg CO2 eq/kg fuel)

Fig. 39 Greenhouse gas (GHG) emissions during production of various fuels.


Ammonia Production 71

The production of the different fuels is compared in terms of abiotic depletion of sources as shown in Fig. 40. Ammonia fuel
has the lowest abiotic depletion value compared to others although the production process may be fossil fuel-based. There are
multiple pathways for ammonia production. Ammonia is cleaner when produced from renewable resources. Figs. 41 and 42
compare the environmental impacts of various ammonia production pathways. Hence, ammonia from renewable resources has
the least environmental impact. Furthermore, ammonia from hydrocarbon cracking and underground coal gasification is most
environmentally benign option among conventional methods.
When renewable sources-based ammonia production options are compared as shown in Fig. 42, ammonia from PV and
biomass-based electrolysis routes yield higher global warming values although they are quite lower than conventional methods.
The abiotic resources are natural resources including energy resources, such as iron ore and crude oil, which are considered as
nonliving. The abiotic depletion is highest for coal electrolysis-based ammonia production methods and followed by heavy oil and
natural gas-based methods as it is illustrated in Fig. 43. This is due to the fact that coal, heavy oil, and natural gas are the
primary source of energy and feed source as well, it indicates the massive consumption of fossil fuels for a unit mass of ammonia
produced.
The impact on human health due to human toxicity is maximum for the ammonia production from coal and heavy oil-based
electrolysis methods where the maximum is to 2.92 kg 1,4-DB-eq/kg of ammonia for coal-based electrolysis method. Ammonia
from both underground coal gasification and naphtha cracking-based methods yield lowest human toxicity values as seen in
Fig. 44 among conventional option. On the other hand, renewable-based ammonia production methods, such as tidal&waves,
municipal waste, geothermal and biomass have lower toxicity values.

Ammonia (from wind electrolysis)


Ammonia (from PV electrolysis)
Ammonia (hydrocarbon cracking)
Ammonia, steam reforming, liquid, at plant
Natural gas, at production
Ammonia, partial oxidation, liquid, at plant
Natural gas, liquefied, at liquefaction plant
Naphtha, at refinery
Diesel, low-sulfur, at refinery
Propane/ butane, at refinery
Petrol, unleaded, at refinery
0 0.005 0.01 0.015 0.02 0.025 0.03
3
Abiotic depletion kg Sb eq/kg (m for natural gas)
Fig. 40 Abiotic depletion values during production of various fuels.

Naphta cracking

Underground coal gasification with carbon capture

Nuclear high temperature electrolysis

Photovoltaic electrolysis

Steam methane reforming

Partial oxidation of heavy oil

Coal gasification

Nuclear 3 step Cu−CI cycle

Natural gas electrolysis

Heavy oil electrolysis

Coal electrolysis

0 2 4 6 8 10 12 14 16
Global warming potential (kg CO2 eq/kg ammonia)

Fig. 41 Global warming values of various conventional ammonia production methods.


72 Ammonia Production

Ammonia from tidal and wave electrolysis


Ammonia from municipal waste electrolysis
Ammonia from geothermal electrolysis
Ammonia from DG biomass
Ammonia from CFBG biomass
Ammonia from hydropower electrolysis
Ammonia from hydropower (river) electrolysis
Ammonia from hydropower (reservoir) electrolysis
Ammonia from wind electrolysis
Ammonia from biomass electrolysis
Ammonia from PV electrolysis
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Global warming 500a (kg CO2 eq/kg ammonia)

Fig. 42 Global warming values of renewable ammonia production methods. CFBG, circulating fluidized bed gasification.

Ammonia from tidal and wave electrolysis


Ammonia from municipal waste electrolysis
Ammonia from geothermal electrolysis
Ammonia from DG biomass
Ammonia from CFBG biomass
Ammonia from coal gasification
Ammonia from UCG without CCS
Ammonia from UCG with CCS
Ammonia from biomass electrolysis
Ammonia from hydropower electrolysis
Ammonia from hydropower (river) electrolysis
Ammonia from hydropower (reservoir) electrolysis
Ammonia from wind electrolysis
Ammonia from nuclear electrolysis
Ammonia from PV electrolysis
Ammonia from nuclear high temperature electrolysis
Ammonia from naphta cracking
Ammonia from SMR
Ammonia from partial oxidation of heavy oil
Ammonia from nuclear 4 step Cu−Cl cycle
Ammonia from nuclear 3 step Cu−Cl cycle
Ammonia from nuclear 5 step Cu−Cl cycle
Ammonia from natural gas electrolysis
Ammonia from heavy oil electrolysis
Ammonia from bituminous coal electrolysis
Ammonia from coal electrolysis
0 0.02 0.04 0.06 0.08 0.1 0.12
Abiotic depletion (kg Sb eq./kg ammonia)
Fig. 43 Abiotic depletion values of various conventional and renewable ammonia production methods. CCS, carbon capture storage; CFBG,
circulating fluidized bed gasification; SMR, Steam methane reforming; UCG, underground coal gasification.

The Eco-Indicator 99 method combines emissions a single score at the highest level. Similar to other results, the single score of
ammonia production from coal electrolysis-based methods have the maximum values while coal gasification methods have the
minimum values among other methods. As illustrated in Fig. 45, heavy oil-based ammonia production method has the highest
environmental impact on resources. The lowest single score is calculated for UCG with carbon capture storage (CCS) method
which can be evaluated as the most environmentally benign method among conventional methods.
Ammonia Production 73

Ammonia from tidal and wave electrolysis


Ammonia from municipal waste electrolysis
Ammonia from geothermal electrolysis
Ammonia from DG biomass
Ammonia from CFBG biomass
Ammonia from naphta cracking
Ammonia from coal gasification
Ammonia from UCG without CCS
Ammonia from UCG with CCS
Ammonia from biomass electrolysis
Ammonia from hydropower electrolysis
Ammonia from hydropower (river) electrolysis
Ammonia from hydropower (reservoir) electrolysis
Ammonia from SMR
Ammonia from wind electrolysis
Ammonia from nuclear 5 step Cu−Cl cycle
Ammonia from nuclear 4 step Cu−Cl cycle
Ammonia from nuclear 3 step Cu−Cl cycle
Ammonia from PV electrolysis
Ammonia from nuclear high temperature electrolysis
Ammonia from natural gas electrolysis
Ammonia from bituminous coal electrolysis
Ammonia from nuclear electrolysis
Ammonia from partial oxidation of heavy oil
Ammonia from heavy oil electrolysis
Ammonia from coal electrolysis
0 0.5 1 1.5 2 2.5 3 3.5

Human toxicity 500a (kg 1,4-DB eq/kg ammonia)


Fig. 44 Human toxicity values of various ammonia production methods. CCS, carbon capture storage; CFBG, circulating fluidized bed gasification;
SMR, steam methane reforming; UCG, underground coal gasification.

1.0
Human health Ecosystem quality Resources
0.9
0.8
0.7
0.6
Rank

0.5
0.4
0.3
0.2
0.1
0.0
Ammonia from coal electrolysis
Ammonia from bituminous coal
electroysis
Ammonia from heavy oil electrolysis

Ammonia from Natural gas electrolysis


Ammaonia from partial oxidation of
heavy oil
Ammonia from nuclear 3 step Cu−CI cycle

Ammonia from nuclear 5 step Cu−CI cycle

Ammonia from nuclear 4 step Cu−CI cycle

Ammonia from nuclear electrolysis


Ammonia from nuclear high temperature
electrolysis
Ammonia from SMR

Ammonia from UCG without CCS

Ammonia from coal gasification

Ammonia from naphta cracking

Ammonia from UCG with CCS

Fig. 45 Single score values of conventional ammonia production methods according to Eco-Indicator 99. CCS, carbon capture storage; SMR,
steam methane reforming; UCG, underground coal gasification.
74 Ammonia Production

3.2.4.3 Case Study 2 Conclusions


Ammonia is a carbon-free chemical energy carrier suitable for use as a transportation fuel. Furthermore, ammonia has a high
octane rating (110–130), can be thermally cracked to produce hydrogen fuel using only approximately 12% of the HHV. It has a
well-established production and distribution infrastructure and has zero GWP. In addition to its attractive qualities as a fuel,
ammonia is widely used as a NOx reducing agent for combustion exhaust gasses using selective catalytic reduction (SCR), and its
capacity as a refrigerant can be applied to recover and further utilize engine heat that would otherwise be lost. In terms of
environmental sustainability, ammonia can be produced using either fossil fuels, or any renewable energy source, using heat
and/or electricity, which allows for the evolution of ammonia production methods and technologies in parallel with sustainable
development. The following concluding remarks are expressed based on the current study:
• In terms of human toxicity, coal, and heavy oil-fired power plant-based electrolysis methods for ammonia production have
highest values.
• Tidal&Waves, municipal waste, and geothermal-based ammonia production routes have lowest abiotic depletion, global
warming, and human toxicity values respectively among all methods.
• Nuclear electrolysis and naphtha cracking-based ammonia production methods have least effect on climate change among
conventional methods while Tidal&Waves method is the most environmentally benign method in terms of climate change and
global warming.
• The renewable sources with their improved efficiency can reduce the overall environmental footprint and can replace the
current fossil fuel-based centralized ammonia production facilities.
• As the cost of renewable electricity catches the level of conventional electricity, renewable energy-based ammonia production
systems will continue to gain practicality and popularity.

3.2.5 Case Study 3: Comparative Ammonia Production, Storage and Transport Scenarios

In this section, various case studies ranging from high-pressure electrolysis process to thermal methane plasma disassociation are
comparatively studied. Possible improvements in Haber–Bosch process are investigated by eliminating the main feed gas com-
pressor and utilizing the excess heat in the reactor. The feasibility study of the selected options is conducted.

3.2.5.1 Case 1: Ammonia Production via High-Pressure Electrolysis


The average cost of ammonia production from the electrolysis-based systems are approximately 20–25% of hydrogen production
cost as previously given in Ref. [7] for various ammonia production methods. 17.8% of ammonia is hydrogen in weight, and
around 3% of ammonia production cost comes from air separation-based nitrogen production. In this case study, the lower limit
has been taken to calculate the cost of ammonia production from high-pressure electrolysis-based systems. Therefore, ammonia
production costs are calculated as the 20% of hydrogen production cost. The schematic diagram of high-pressure electrolysis-based
ammonia production system is given in Figs. 46 and 47. In this case, hydrogen compressor is eliminated by using pressurized water

Air
Low cost hydropower

Hydroelectric
power plant

Electricity
Electricity
High pressure
water High pressure
Cryogenic air
PEM O2 separation
electrolyzer O2

Pump

Compressor

Water High pressure Haber−Bosch High pressure


H2 ammonia N2
synthesis
(no main compressor)

NH3

Ammonia
storage

Fig. 46 Schematic diagram of high-pressure electrolysis-based ammonia production system.


Ammonia Production 75

Hydropower

Water
Hydroelectric
power plant

High pressure
Nitrogen (N2) water Electricity

High pressure electrolysis


Haber−Bosch with no main compressor
ammonia synthesis

Ammonia
(NH3)

Fig. 47 Schematic diagram of energy and material flows of high-pressure electrolysis-based ammonia production system.
High pressure electrolyzer current density

5 cent/kWh
2787
3.5 cent/kWh
(mA/cm2)

5 cent/kWh
1858
3.5 cent/kWh

5 cent/kWh
929
3.5 cent/kWh

0.00 0.20 0.40 0.60 0.80 1.00 1.20

Ammonia production cost ($/kg)

Total ammonia cost ($/kg) Total ammonia cost ($/kg)


(Electricity : US 5 cent/kWh) (Electricity : US 3.5 cent/kWh)
Fig. 48 Cost comparison of ammonia production from high-pressure polymer electrolyte membrane (PEM) electrolysis and Haber–Bosch plant
based on various electricity price and current electrolyzer densities at 345 bar. Data from Cropley C, Norman T. Low-cost, high-pressure hydrogen
generator. Final report on DOE cooperative agreement No. DE-FG36-04GOI3029; 2008.

and high-pressure electrolysis system. However, the nitrogen is used as a gaseous form and still need to be compressed by the
compressor. The amount of work decreases substantially because nitrogen constitutes about 82% of the feed gases. The required
electricity is supplied via low-cost hydroelectric plant to the electrolyzer, cryogenic air separation unit, external nitrogen com-
pressor, pump and recycling compressor inside the Haber–Bosch process.
Using the data from the previous chapter given in high-pressure electrolysis hydrogen production prices, the costs of ammonia
production via high-pressure electrolysis-based electrolysis at two different electricity prices are illustrated in Fig. 48.

3.2.5.2 Case 2: Ammonia Production via Liquid Nitrogen and High-Pressure Electrolysis
In this case, opportunities of utilizing the excess heat of Haber–Bosch process are investigated as illustrated in Figs. 49 and 50.
The negative value of enthalpy in the ammonia synthesis process indicates that the reaction is exothermic, releasing approximately
76 Ammonia Production

Air
Low cost hydropower

Hydroelectric
power plant

Electricity
Electricity

High pressure
water High pressure
Cryogenic air
PEM O2 separation
electrolyzer O2

Pump
Cryogenic
pump
Excess heat
High pressure
Water H2 Haber−Bosch High pressure
ammonia liquid N2
Evaporator
synthesis
(no main compressor)

NH3

Ammonia
storage

Fig. 49 Schematic diagram of high-pressure electrolysis and liquid nitrogen pumping-based ammonia production system. PEM, polymer electrolyte
membrane.

Low cost hydropower

Water
Hydroelectric
power plant

High pressure
water Electricity
Liquid
nitrogen (N2)
High pressure electrolysis
Evaporator Haber−Bosch with no main compressor
ammonia synthesis

Excess heat

Ammonia
(NH3)

Fig. 50 A diagram of energy and material flows of high-pressure electrolysis and liquid nitrogen pumping-based ammonia production system.

2.7 GJ/t NH3 heat in ammonia production. This is equivalent to about 8% of the energy input for the entire process. It means that
_ HB;Specific ¼ 2700 kJ=kg ammonia. For the ammonia production facility with a capacity of 300 t/day,
heat dissipation is around Q
the amount of required nitrogen mass flow rate is 246.7 t/day which corresponds to about 2.90 kg/s flow rate.
Ammonia Production 77

In this case, hydrogen is produced via high-pressure electrolysis and sent to Haber–Bosch reactor. Cryogenic air separation
produces nitrogen as the liquid end product. The liquefaction of nitrogen process will require more energy compared to the
gaseous end product. The liquid nitrogen is pumped to reaction pressure about 200 bar by the cryogenic pump.
The excess heat in Haber–Bosch reactor is utilized for the vaporization of high-pressure liquid nitrogen to obtain high-
pressure gaseous nitrogen for ammonia synthesis reaction. The required electricity is supplied via low-cost hydroelectric power
plant to the electrolyzer, cryogenic air separation unit, pumps and recycling compressor inside the Haber–Bosch process.
An air separation facility with a capacity of around 250 t/day is considered. The outlet pressure of air separation unit is
generally at 8 bar. The liquid nitrogen at 8 bar is pressurized using cryogenic pumps up to 200 bar.
m _ Pump ¼ m
_ N2  hliquid nitrogen 1 þ W _ N2  hliquid nitrogen 2 ð8Þ
The required pump power is calculated as 52.36 kW. The liquid nitrogen enters to a vaporizer which utilizes the excess heat in
Haber–Bosch process. The outlet temperature of vaporized unit is set to 251C. The mass flow rates are calculated as
m_ ammonia ¼ 300  243600
1000 _ hydrogen ¼ 53:3  243600
¼ 3:47 kg=s and m 1000
¼ 0:61 kg=s.
Considering the 300 t/day ammonia plant, the Haber–Bosch reactor releases 9375 kW heat. The hydrogen and nitrogen gas
mixture are sent to the reactor. The reaction temperature for Haber–Bosch process is approximately 4501C. The temperature of
hydrogen is assumed as 1001C after high-pressure electrolysis. The required heat of vaporization for nitrogen at T2 ¼  1801C is
calculated as 506.1 kW.
_ Haber Bosch ¼ Dhvaporization;liquid nitrogen  m
Q _ N2 ð9Þ
Besides, the mixture gasses hydrogen and nitrogen require 4589 kW heat in order to reach the reaction temperature.
_ Hydrogen ¼ m
Q _ H2  ðhhydrogen 2  hhydrogen 1 Þ ð10Þ

_ Nitrogen ¼ m
Q _ N2  ðhnitrogen 2  hnitrogen 1 Þ ð11Þ
where the initial temperature of hydrogen is 1001C and initial temperature of nitrogen is 251C, and final temperature is 4501C.
_ Excess heat ¼ Q
Q _ HaberBosch  Q
_ Reaction temperature ð12Þ
Therefore, 4786 kW excess heat is available which can be utilized in vaporization of liquid nitrogen. Finally, the amount of
excess heat from Haber–Bosch reactor is enough for the vaporization of liquid nitrogen. On the other hand, if gaseous nitrogen is
pressurized to the reaction temperature instead of liquid pumping, the required compressor power to have compressed nitrogen
from 8 to 200 bar is calculated to be 1057 kW. In total, liquid nitrogen pumping and vaporization require around 559 kW power.
This indicates that there is around 52% reduction in energy requirement by using liquid nitrogen pumping and vaporization
obtained by the excess reactor heat. The cost of 5 centrifugal compressors in the 300 t/day ammonia synthesis synloop represents
50% of the overall installed cost of synthesis loop. When these compressors are eliminated, nearly $11 million capital cost
reduction can be achieved. The required power for 300 t/day ammonia plant is calculated as 145 MW. Around 7.7 MW of this
power corresponds to synloop compressors. Therefore, the required power can decrease 5.3% for the overall plant. Producing
ammonia directly from high-pressure electrolysis near the ammonia production facility eliminates the compression, storage and
delivery processes of hydrogen which brings down the overall ammonia production cost.

3.2.5.3 Ammonia Fueled Generators for Stand-Alone Power Production by Transporting Ammonia via Tanker Trucks
Diesel engines are simply compression-ignition engines and can operate on a variety of different fuels, depending on configuration
and location. Where a gas grid connection is available, gas is often used, as the gas grid will always remain pressurized even during
almost all power cuts; in more rural situations, or for low load factor plant, diesel fuel derived from crude oil is a common fuel; it
is less likely to freeze than heavier oils.
Ammonia can also be used in the following applications in remote communities:

• Fertilizers in the agriculture.


• Fuel for farm tractors.
• Heavy trucks.
• Vessels.
• Heavy duty engine fuel.
• Off peak and excess power storage and distributed power generation.

Some of the advantages of using ammonia in remote communities are listed as follows:

• Ammonia is currently second largely synthesized chemical in the world. It is already being transported in large quantities using
tanker trucks.
• Ammonia can be utilized as dual fuel with diesel, propane, hydrogen, etc.
• Power transmission line is eliminated by using distributed power generation.
• Zero GHG during utilization in diesel cycle generators.
• Ammonia transportation pipelines can be installed which can also be used for natural gas, propane, etc.
78 Ammonia Production

• Using solar energy, ammonia can be used for combined heat and power production.
• Reduced use of diesel for electricity generation will also have significant effects on the environmental impact of electricity in the
communities. It will improve environmental quality within the communities by limiting diesel fuel spills, pollution resulting
from combustion and noise pollution of the diesel system. Diesel generation creates significant emissions in remote com-
munities, causing local pollution, and GHG releases.
On the other hand, there might be some disadvantages depending on the fuel transportation requirement.

• Transportation of ammonia using tanker trucks may bring additional costs and maintenance requirements for the roads.
• Since the LHV of ammonia is lesser than diesel, the transportation requirement is higher.
• Without power transmission lines, power generation for the provincial grid would not be possible which can provide a revenue
source for communities.

3.2.5.4 Renewable Energy-Based On-Site Ammonia Production and Utilization in Ammonia Fueled Generators for Stand-
Alone Power Production
Any renewable energy source can be used for on-site ammonia production. Northern Ontario in Canada has rich hydropower,
wind and solar energy sources. In order to compensate the intermittency problem of solar and wind energy, an integrated system
coupling these renewable resources can be simply utilized.

Advantages:

• Ammonia can be produced from renewable energy resources on-site. Canada has more than 10 large-scale ammonia pro-
duction facilities. Currently, SMR method is utilized for ammonia production. However, water can be dissociated into
hydrogen and oxygen using renewable energy and then produced hydrogen can be combined with nitrogen from the air to
produce ammonia.
• Power transmission line is eliminated by using distributed power generation.
• Zero GHG during utilization in diesel cycle generators.
• There are also a number of potential hydroelectric sites in potential size up to about 30 MW which can be utilized for on-site
ammonia production.
• Ammonia fuel may provide an alternative to electricity for transmission, annual-scale firming storage, and energy supply
integration.
• Converting stranded, curtailed, or spilled renewable energy source electricity, at the sources, to ammonia fuel, allows harvest,
transmission, and storage of this stranded renewable energy, for a degree of community energy independence.

Disadvantages:

• Usage of only one renewable resource may cause storage problems because of intermittency.
• Sitting of ammonia storage tank or tank farm for preferably at a distance downwind of the community may be needed to
prevent any large accidental ammonia leak.

3.2.5.5 Cost Comparisons


In this section, alternative diesel generator fuels are comparatively assessed considering the fuel costs in Table 4 in terms of annual
and lifetime (40 years) period.
For a 2000 kW diesel generator, the parameters listed in Table 5 are considered for cost analyses.
As seen in Table 5 and Fig. 51, the operation of diesel generators fueled with ammonia yields the lowest cost.
These results are for a specific fleet operating within primarily within the greater Toronto area (GTA) and for 16,000 average
kilogram of payload capacity. The cost terms are in Canadian dollar (CAD) here. The congestion premium noted would obviously
be less for fleets that operate for longer periods of time in non-congested traffic areas. The truck transportation cost values shown
in Table 6 are taken from the report of “Estimation of costs of heavy vehicle use per vehicle-kilometer in Canada” by transport
Canada economic analysis directorate [41]. Accordingly, the fuel and transportation requirements are listed in Table 7.
If 1500 km distance is considered for fuel transportation using tractor–trailer type trucks, the calculations listed in Table 8 are
obtained.

Table 4 Current market prices


of ammonia, diesel and hydrogen

Fuel Cost ($/kg)

Ammonia 0.28
Hydrogen 3.2
Diesel 0.85
Ammonia Production 79

Table 5 Comparison of diesel generator operation for various fuels

Fuel Lower heating value Fuel mass flow rate in Unit cost Daily cost Yearly Life time cost ($)
(LHV) (kJ/kg) diesel engine (kg/s) ($/kg) ($/day) (24 h) cost ($) (40 years)

Ammonia 18,646 0.32 0.28 7741.44 2,825,625 113,025,024


Hydrogen 120,210 0.05 3.2 13,824 5,045,760 201,830,400
Diesel 42,791 0.1396 0.85 10252.22 3,742,061 149,682,470

202
200
Life time cost of operation (million $)

150
150

113

100

50

0
Ammonia Diesel Hydrogen
Fig. 51 Life time (40 years) cost comparison of diesel generator operations driven by various fuels.

Table 6 Estimated per-kilometer truck cost for sample fleet in greater Toronto area (GTA)

Truck type Per kilometer costs Congestion premium

With congestion Without congestion

Straight $2.22 $1.89 17%


Tractor–trailer $2.68 $2.34 14.5%

Table 7 Fuel and transportation requirements

Fuel Fuel mass flow rate in Required fuel mass Required fuel mass Required number of Required number of
diesel engine (kg/s) in 1 year (t) in 40 years (t) truck transportation truck transportation
in 1 year in 40 years

Ammonia 0.32 10,091 403,660 631 25,229


Hydrogen 0.05 1,576 63,072 99 3,942
Diesel 0.1396 4,402 176,097 276 11,007

Table 8 Cost of transportation for various fuels

Fuel Number of truck Number of truck Transportation Cost of transportation Life time cost of
transportation in transportation in cost ($/km) ($/year) transportation
1 year 40 years ($/40 years)

Ammonia 631 25,229 2.4375 2,307,093 92,243,531


Hydrogen 99 3,942 2.4375 361,968 14,412,937
Diesel 276 11,007 2.4375 1,009,125 40,244,343
80 Ammonia Production

Ammonia, hydropower electrolysis, at plant

Diesel, low-sulfur, at refinery

Petrol, unleaded, at refinery

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7


Global warming 500a (kg CO2 eq/kg)

Fig. 52 Greenhouse gas (GHG) emissions during production of 1 kg petrol, diesel, and ammonia.

The life time cost of ammonia transportation is calculated more than diesel because of lower LHV and more mass requirement.
However, the major advantage of ammonia is that it can be produced on-site which eliminates the transportation costs. Never-
theless diesel cannot be generated on-site for utilization in remote regions. In addition, there are much more alternative ways for
power production using ammonia such as gas turbines.
Fig. 52 is obtained via CML 2001 environmental impact assessment method. In this method, the results are given in terms of
CO2 equivalent. The production of ammonia from hydropower, which is an abundant source of renewable energy in remote
communities, yields the lowest environmental impact. Here, ammonia is produced on-site whereas diesel and petrol are produced
at refinery. Therefore, additional GHG emissions will be caused by the transportation of these petroleum fuels. This means that
on-site production of ammonia will be a more environmentally benign option instead of diesel production and transportation.

3.2.5.6 Case Study 3 Conclusions


The following remarks are noted based on the calculations and investigations.

• As the global warming and energy issues become important topics to be considered for the future of remote communities,
alternative solutions for power and heat generation need to be investigated.
• Currently, diesel generators running on diesel are utilized for Northwestern Ontario remote communities. This is one of the
most expensive and environmentally damaging options.
• The significance of distributed electricity generation is emphasized in many studies throughout the world. Hence, producing
power and heat via stand-alone facilities are being encouraged by the governments and decision makers.
• As an alternative and sustainable fuel, ammonia, can be utilized in the diesel generators by minor modifications. Ammonia has
no GHG emission during utilization in the diesel generators. Hence, it is one of the most environmentally benign fuel among
other alternatives.
• Transportation and storage of ammonia are already available and well-known since it is the second largest produced chemical
in the world. This implies that instead of diesel, ammonia can easily be transported and stored.
• Ammonia can be produced on-site using renewable energy resources, such as wind, solar and hydropower which are already
available in these remote communities. This brings minimum transportation cost. However, diesel needs to be transported for
long distances.
• The operation of diesel engines with fuel ammonia has the lowest cost based on the current market prices. Considering the
technology development of ammonia production, the cost of ammonia will continue to decrease which will bring additional
reductions in total cost.
• Production and utilization of ammonia in diesel generators have significantly lower environmental impacts in terms of climate
change and global warming.
• As a result, ammonia usage in remote communities for power and heat production will bring significant cost and environ-
mental benefits.

3.2.6 Case Study 4: Ammonia Production Scenarios in Canada

In this case study, various ammonia production routes in Canada are investigated by considering the production, storage, and
transport phases.

3.2.6.1 Ammonia Production Using Water Electrolysis From Low-Cost Hydropower and Wind Energy
The system modeled is a standalone grid powered PEM electrolyzer system with a total hydrogen production capacity of
50,000 kg/day [42]. This corresponds to about 280 t/day ammonia production plant. The system is based on a generic system
Ammonia Production 81

Table 9 Analysis parameters for electrolysis-based methods

Energy efficiency calculations


H2 outlet pressure psi 1000
Stack electrical usage
Cell voltage Volts/cell 1.6625
Voltage efficiency % Lower heating value (LHV) 74.0%
Dryer loss % of gross H2 1.5%
Permeation Loss % of gross H2 2.0%
Total stack energy usage per mass kWhe/kgNet H2 46.67
net H2
BOP loads
Power inverter efficiency % 97%
Inverter electrical load kWhe/kgNet H2 1.44
Dryer thermal load kWhtkgNet H2 0.31
Dryer efficiency kWhe/kWhtherm 3.30
Dryer electrical load kWhe/kgNet H2 1.02
Misc electrical load kWhe/kgNet H2 1.08
Total BOP electrical load kWhe/kgNet H2 3.54
Summary
Stack electrical usage kWhe/kg H2 46.67
BOP electrical usage kWhe/kg H2 3.54

Total system electrical usage per mass net H2 kWhe/kgNet H2 50.2


Total system electrical usage per mass net H2 (high-pressure electrolysis) kWhe/kgNet H2 40.6
Reference year 2007
Assumed start-up year 2020
Basis year 2012

Source: DOE Hydrogen and Fuel Cells Program. DOE H2A production analysis. H2A central hydrogen production model version 3.1. Available From:
https://www.hydrogen.energy.gov/h2a_production.html; 2017 [accessed 07.01.17].

using input from several key industry collaborators with commercial experience in PEM electrolysis. The electrolyzer units use
process water and grid electricity for electrolysis. Two different electrolysis process namely high-pressure electrolysis and standard
electrolysis are considered in the assessments where the parameters are given in Table 9.
The required power for 300 t/day ammonia plant was calculated as 145 MW. Around 7.7 MW of this power corresponds to
synloop compressors. Therefore, if high-pressure electrolysis is utilized, the required power will decrease 5.3% for the overall plant.
The average cost of ammonia production from the electrolysis-based systems are approximately 20–25% of hydrogen production
cost as previously given in Ref. [7] for various ammonia production methods. In the current analyses, ammonia production costs
are calculated as the 20% of hydrogen production cost. The source of electricity is taken as wind energy and hydroelectric power
plant. These two renewable sources are significant and dominant in various locations of Canada. In the case studies, hydropower
plant is located in Ontario whereas wind power plants are located in Newfoundland and Labrador. The province has a high
potential of wind energy as compared in the Fig. 53 and Table 10. The average wind capacity factor for Newfoundland is taken to
be as 40% [43].
The hydroelectric and wind energy-based electricity production costs are taken as shown in Table 11.
A project is on the way by Siemens for green ammonia produced by wind energy. They plan to utilize the excess wind power in
ammonia production by investing in wind-powered electrolysis for hydrogen production and then ammonia synthesis. They
propose it as a fuel blending for vehicles. The building of a 30 kW trial with an approximate cost of $2.8 million will start. Since
the electrolysis cost is still in the process of decreasing, they presume to have ammonia commodity in less than 10 years [44].

3.2.6.2 Ammonia Production From Steam Methane Reforming With CO2 Capture and Sequestration
Since SMR is currently the primary production route of ammonia, a comparative assessment can give valuable insights into
alternative methods. In this method, natural gas is fed to the plant from the pipeline at a pressure of 31 bar. The desulfurized
natural gas feedstock is mixed with process steam to be reacted over a nickel-based catalyst contained inside of a system of high
alloy steel tubes. The reforming reaction is strongly endothermic, and the metallurgy of the tubes usually limits the reaction
temperature to 760–9261C. The flue gas path of the fired reformer is integrated with additional boiler surfaces to produce steam.
A portion of this steam is superheated to 31 bar and 3991C, to be added to the incoming natural gas. Additional steam from the
boiler is used to regenerate the CO2. After the reformer, the process gas mixture of CO and H2 passes through a heat recovery step
and is fed into a water gas shift reactor to produce additional H2. Two different locations are chosen for SMR-based ammonia
production options namely Edmonton, Alberta and Toronto, Ontario. The following grid electricity prices are taken into account
in the analyses as shown in Fig. 54. The assumptions and design parameters are given in Table 12.
82 Ammonia Production

14
3%/yr
6%/yr
12 9%/yr

Cost of electricity (cents/kWh)


12%/yr
10

0
er

io

c
ta

I
ick

nd
PE
be
in

pe

ar
uv

er

la
w
eg

ue

S+
ni

nt
b

nd
ns
co

Al

in

N
u
n

u
W

S.
S.

Br
Va

fo
S.

ew
ew

N
N
Fig. 53 Cost of wind energy-based electricity with rates of return on the capital investment of 3 to 12%/ yr for the base case. Modified from Harvey
L.D.D. The potential of wind energy to largely displace existing Canadian fossil fuel and nuclear electricity generation. Energy 2013;50:93–102.

Table 10 Comparison of cost of wind energy-based electricity for various locations in Canada

Location Capital cost of Transmission Capital cost of Capacity Cost of Cost of


wind turbine distance (km) transmission factor transmission electricity
($/kW) line ($/kW) (cents/kWh) (cents/kWh)

Vancouver 1862 677 839 0.460 1.11 5.29


South Alberta 2059 365 683 0.402 1.02 6.03
Regina 2165 2043 1521 0.506 1.86 6.39
Winnipeg 1593 3401 2200 0.527 2.54 6.11
S. Ontario 2083 1820 1410 0.531 1.61 5.79
S. Quebec 1952 1465 1232 0.512 1.46 5.49
New Brunswick 1872 832 933 0.492 1.15 5.14
Nova Scotia 2014 831 916 0.529 1.05 5.04
Newfoundland 1648 87 544 0.458 0.72 4.50

Source: Data from Harvey LDD. The potential of wind energy to largely displace existing Canadian fossil fuel and nuclear electricity generation. Energy 2013;50:93–102.

Table 11 The electricity prices taken for the analyses of electrolysis-based ammonia production

Method Province Electricity price ($/kWh)

Low-cost hydropower high-pressure electrolysis Ontario 0.0275


Low-cost hydropower electrolysis Ontario 0.0275
Wind high-pressure electrolysis Newfoundland 0.045
Wind electrolysis Newfoundland 0.045

3.2.6.2.1 Current ammonia retail prices


The ammonia prices have fallen to the lowest levels since April 2009. Ammonia prices fell primarily due to lower natural gas costs.
There are positives and negatives for falling ammonia prices. Since most of the ammonia is upgraded to other fertilizers, lower
prices benefit companies that do not sell it directly [45]. The following Figs. 55 and 56 show the retail fertilizer prices in
comparison with ammonia in the market. The currency is dollar ($). Ammonia has the highest prices among other fertilizers.

3.2.6.3 Cost Analyses Results


Based on the previously mentioned assumptions and calculations, the following ammonia production costs are obtained for
various scenarios in Canada as shown in Table 13. As comparatively shown in Table 13 and Fig. 57, the highest ammonia cost is
Ammonia Production 83

10.02 Halifax, NS
8.9 Charlottetown, PE
7.14 Moncton, NB
6.55 Regina, SK
6.13 Ottowa, ON
5.84 Vancouver, BC
5.55 Toronto, ON
4.9
Montreal, QC
4.77
4.74 St. John’s, NL
4.22 Calgary, AB
4.02 Edmonton, AB
Winnipeg, MB
0 2 4 6 8 10
Average electricity price (Canadian cent/kWh)
Fig. 54 Comparative index of electricity prices for large power customers with a monthly consumption of 3,060,000 kWh and a power demand
of 5000 kW. Data from Hydro-Québec. Comparison of Electricity Prices in Major North American Cities. Available From: http://www.hydroquebec.
com/publications/en/corporate-documents/comparaison-electricity-prices.html; 2017 [accessed 07.01.17].

Table 12 Analysis parameters for steam methane reforming (SMR)


method

Energy efficiency of SMR 0.82


Energy efficiency of shift reactor 0.95
Energy efficiency of PSA hydrogen separation 0.8
Reference year 2007
Assumed start-up year 2020
Basis year 2009
Usage (kWh/kg H2) 0.6
Pressure (bar) 20

Source: DOE Hydrogen and Fuel Cells Program. DOE H2A production analysis. H2A central
hydrogen production model version 3.1. Available From: https://www.hydrogen.energy.gov/
h2a_production.html; 2017 [accessed 07.01.17].

DAP Potash Urea Ammonia

$700
$650
$600
per ton

$550
$500
$450
$400
$350
13

13

14

14

14

14

15

15

15

15

16
/

/
09

09

09

09

09

09

09

09

09

09

09
/

/
09

12

03

06

09

12

03

06

09

12

03

Fig. 55 Average retail fertilizer prices in the United States. Data from Weekly Fertilizer Review. Ammonia prices. Available From: http://www.
farmfutures.com; 2017 [accessed 07.01.17].

calculated for wind electrolysis route in Newfoundland. However, by using high-pressure electrolysis the cost of ammonia can be
decreased down to 0.46 $ which is quite closer to conventional SMR method. The lowest cost is observed in hydrocarbon
dissociation based on the price given in Ref. [46].
The hydrogen production costs are calculated using H2A Central Hydrogen Production Model, Version 3.1 of U.S. Department of
energy (DOE) [42]. On the other hand, the cost of SMR with CCS-based ammonia is slightly lower in Alberta because of lower
electricity prices and distance to the natural gas wells.

3.2.6.3.1 Storage cost


Besides the production costs, various scenarios for storage and transportation of ammonia are developed in this section. The
storage of ammonia requires not only a capital cost for the facility but also an operating energy cost depending on if pressurized or
84 Ammonia Production

Ammonia DAP Potash Urea


$1200

$1000

$800

per ton
$600

$400

$200
Sep-08 Sep-09 Sep-10 Sep-11 Sep-12 Sep-13 Sep-14 Sep-15
Fig. 56 Illinois fertilizer prices including ammonia since 2008. Data from Weekly Fertilizer Review. Ammonia prices. Available From: http://www.
farmfutures.com; 2017 [accessed 07.01.17].

Table 13 Cost of ammonia production for different scenarios in Canada

Method Province Electricity price Hydrogen cost Ammonia cost Reference


($/kWh) ($/kg) ($/kg)

Low-cost hydropower high-pressure Ontario 0.0275 1.73 0.3287 [42]


electrolysis
Low-cost hydropower electrolysis Ontario 0.0275 2.12 0.424 [42]
Wind high-pressure electrolysis Newfoundland 0.045 2.44 0.4636 [42]
Wind electrolysis Newfoundland 0.045 3 0.6 [42]
Hydrocarbon dissociation Alberta – 1.5 0.3 [46]
Steam methane reforming (SMR) with Ontario 0.05 2.2 0.44 [42]
carbon capture storage (CCS)
SMR with CCS Alberta 0.042 2.1 0.42 [42]

Steam methane reforming with CO2 capture


and sequestration
Steam methane reforming with CO2 capture
and sequestration

Hydrocarbon dissociation

Wind electrolysis

Wind high pressure electrolysis

Low-cost hydropower electrolysis

Low-cost hydropower high pressure


electrolysis
0 0.5 1 1.5 2 2.5 3
Cost of production ($/kg)

Ammonia cost ($/kg) Hydrogen cost ($/kg)


Fig. 57 Comparison of production cost of ammonia and hydrogen using various routes.

low-temperature storage is used. The energy use and efficiency of low-temperature storage was analyzed for both hydrogen and
ammonia transportation fuels in Ref. [7]. For large-scale storage of ammonia or hydrogen, low-temperature storage is typically
used based on cost considerations. For example, both ammonia and hydrogen can be stored as a liquid at atmospheric pressure if a
low enough temperature is maintained.
The low-temperature storage system consists of a large insulated tank and a refrigeration system to maintain the fuel as a liquid
at the low-temperature. The insulated vessel is only designed with the structural strength to withstand the static pressure of the
Ammonia Production 85

fluid, which greatly reduces the steel content of the vessel compared to pressure storage. Ammonia storage vessels are constructed in a
range of sizes from 4500 to 45,000 t, although typical facilities store between 15,000 and 60,000 t. The ammonia storage capacity
selected for the analyses was 15,000 t, which is the smallest sized commercial facility commonly used by industry. The length of
storage was assumed to be about 6 months, which is based on a storage period between summer and winter seasons.
The storage vessel would be able to store the fuel between seasons to allow for a reliable supply of fuel for vehicles. Efficiency
was defined as the chemical energy stored in the vessel divided by the sum of both the energy input to the system and the chemical
energy stored in the vessel. The storage efficiency calculated for ammonia was 93.6%. The ammonia also enters the liquefaction
process from the ammonia synthesis process as a liquid at  251C, and therefore the amount of heat removal required to achieve
the  331C storage temperature is minimal compared to the 2731C decrease in temperature and phase change for hydrogen
liquefaction [7].
The energy cost would vary depending on the amount of fuel stored in the vessel at any given time and how often the tank is
filled and emptied. If electricity costs 0.08 $/kWh and 6 months of storage is used, then the cost of hydrogen and ammonia storage
was calculated as 0.95 and 0.03 $/kg-H2, respectively, ignoring the capital cost. Combining the energy cost with the capital cost
gives the total storage cost for 6 months of storage to be 14.95 $/kg-H2 for hydrogen, and 0.54 $/kg-H2 for ammonia. Therefore,
ammonia has a cost of storage nearly 30 times less than that of hydrogen [7]. The cost of ammonia storage including the capital
cost yields 0.095 $/kg NH3 as listed in Table 14.
Based on the calculations, the total cost of storage for ammonia for a duration of approximately 6 months and 15,000 t is
found to be 1.425 million US dollar.

3.2.6.3.2 Transportation cost


The transportation cost of ammonia is calculated based on the average costs per kilometer where a truck’s payload capacity is 16 t.
Although there are alternative ways for the transport of ammonia such as pipelines, since there is no ammonia pipeline infra-
structure at the moment in Canada, truck transportation is taken into account. However, for Newfoundland case, ocean trans-
portation is also considered for comparison purposes.
The congestion premium noted would obviously be less for fleets that operate for longer periods of time in non-congested
traffic areas. The truck transportation cost values are taken from the report of “Estimation of costs of heavy vehicle use per vehicle-
kilometer in Canada” [41].
In the following calculations in Tables 15 and 16, it is assumed that 16-tonne ammonia can be carried per truck. The
calculations are performed for a total installed power of 20 MW diesel generator running on ammonia. It is assumed that
ammonia is produced in GTA and transported to Northwestern Ontario with a distance of 1900 km.

Table 14 Ammonia storage energy requirements and storage parameters

Total energy input (kJ/kg NH3) 1532


Total mass (tonne NH3) 15,000
Work input (GJ) 4081
Energy out higher heating value (HHV) (GJ) 59,742
Storage temperature (1C)  33
Efficiency HHV 93.6%
Storage cost ($/kg NH3) 0.095

Source: Bartels JR. A feasibility study of implementing an ammonia economy. Ann Arbor: Iowa
State University; 2008.

Table 15 Calculations of required transportation for a 20 MW required power capacity


remote community in Northwestern Ontario

Fuel mass flow rate Required fuel mass in Required number of truck
(t/day) 6 months (t) transportation in 6 months

277 49,860 3117

Table 16 Total cost of ammonia transportation for a 20 MW remote community


in Northwestern Ontario for 6 months

Required number of truck Transportation Cost of transportation


transportation in 6 months cost ($/km) ($/6 months)

3117 2 11,844,600
86 Ammonia Production

It is calculated that unit transportation cost of ammonia per ton kilometer is 0.125 $/t-kilometer (tkm) NH3. Hence, per kilogram
of ammonia, an additional cost of 0.23 $/kg NH3 is required for the transportation when 1900 km distance is considered.
The total cost of ammonia is calculated using the production, storage, and transportation cost of ammonia for various scenarios
as illustrated in Fig. 58.

3.2.6.3.3 Ammonia production in Ontario


If ammonia is produced using the hydroelectric route from water electrolysis in GTA, Ontario, the approximate distance of
transportation is 1900 km as shown in Table 17. The total cost of ammonia is calculated and comparatively shown in Table 18 and
Fig. 59.
Considering the cost of storage in GTA and transportation to Northwestern Ontario, the total cost of ammonia increases to
about 0.75 $/kg for low-cost hydropower electrolysis whereas it is 0.66 $/kg for high-pressure electrolysis option as seen in Fig. 59.

Fig. 58 Selected scenarios for ammonia production in Canada.

Table 17 Unit cost of ammonia transportation for a remote community in Northwestern Ontario
(ammonia production in greater Toronto area (GTA), Ontario)

Transportation cost per Tonne-kilometer ($/tkm) Distance (km) Unit cost of transportation ($/kg)

0.125 1900 0.2375

Table 18 Total cost of ammonia in Northwestern Ontario for various routes (ammonia production in greater Toronto area (GTA), Ontario)

Cost contributions Low-cost hydropower high- Low-cost hydropower Steam methane reforming with CO2
pressure electrolysis in Ontario electrolysis in Ontario capture and sequestration in Ontario

Production cost ($/kg) 0.3287 0.424 0.44


Storage cost ($/kg) 0.095 0.095 0.095
Transportation cost ($/kg) 0.2375 0.2375 0.2375
Total cost of ammonia in 0.6612 0.7565 0.7725
Northwestern Ontario ($/kg)
Ammonia Production 87

Steam methane reforming

Low-cost hydropower electrolysis

Low-cost hydropower high pressure


electrolysis

0.6

0.62

0.64

0.66

0.68

0.7

0.72

0.74

0.76

0.78

0.8
Total cost of ammonia ($/kg)
Fig. 59 Total cost of ammonia in Northwestern Ontario for different routes in Ontario.

Table 19 Unit cost of ammonia transportation for a remote community in Northwestern Ontario
(ammonia production in Newfoundland and Labrador)

Transportation cost per tonne-kilometer ($/tkm) Distance (km) Unit cost of transportation ($/kg)

0.125 4400 0.55

Table 20 Total cost of ammonia in Northwestern Ontario for various scenarios (ammonia production in
Newfoundland and Labrador)

Cost contribution ($/kg) Wind energy-based high-pressure Wind energy-based electrolysis


electrolysis in Newfoundland in Newfoundland

Production cost 0.4636 0.6


Storage cost 0.095 0.095
Transportation cost 0.55 0.55
Total cost of ammonia 1.1086 1.245

If the cost of hydroelectric can be decreased lower than 0.2 $/kWh, it can be the lowest cost ammonia production among other
methods.

3.2.6.3.4 Ammonia production in Newfoundland and Labrador


If ammonia is produced using wind energy route from water electrolysis in Newfoundland, the approximate distance of trans-
portation is 4400 km as shown in Table 19. The total cost of ammonia is calculated and comparatively shown in Table 20 and
Fig. 60.
Considering the cost of storage in Newfoundland and transportation to Northwestern Ontario, the total cost of ammonia
increases to about 1.1 $/kg for high-pressure electrolysis from wind energy route as seen in Fig. 60. Although the cost of production
is as low as SMR, considering the long transportation distance doubles the cost of ammonia in Northwestern Ontario.
From the perspectives of efficiency, productivity, cost-effectiveness, and reduced emissions, the maritime mode of transpor-
tation is better compared to other modes of transportation. For this case, a further study is conducted for an ocean tanker
transportation of ammonia from Newfoundland to Northwestern Ontario. The high efficiency of maritime operations also
contributes to comparatively lower GHG emissions per tonne-km of cargo moved by ships than by other modes of transportation.
This assures the sustainable and viable nature of the commercial maritime activity. Here, it is also important to note that ammonia
can be used in maritime applications for propulsion.
According to [47], the energy use of ocean shipping corresponds to nearly 7% of truck transportation as shown in Fig. 61.
Hence, the cost of ocean transportation for ammonia is considered to be about 10% of truck transportation for Newfoundland
and Labrador case. In this assessment, the unit cost of ammonia transportation can be evaluated as 0.055 $/kg whereas it was
0.55 $/kg for truck transportation. The total cost of ammonia in Northwestern Ontario can decrease down to 0.62 $/kg for wind
energy-based high-pressure electrolysis and 0.75 $/kg for wind power based electrolysis in Newfoundland and Labrador.
88 Ammonia Production

Wind electrolysis in Newfoundland

Wind high pressure electrolysis


in Newfoundland

1 1.05 1.1 1.15 1.2 1.25 1.3


Total cost of ammonia ($/kg)
Fig. 60 Total cost of ammonia in Northwestern Ontario for different routes (ammonia production in Newfoundland and Labrador).

2.0
1.8
1.6
Energy use (kWh/tkm)

1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
Ocean Rail Truck Air
shipping
Fig. 61 Comparative energy use of various transportation types. Data from John P. Energy transportation and tanker safety in Canada. Fraser
Institute, http://www.fraserinstitute.org; 2015.

Table 21 Unit cost of ammonia transportation for a remote community in Northwestern Ontario (ammonia
production in Alberta)

Transportation cost per tonne-kilometer ($/tkm) Distance (km) Unit cost of transportation ($/kg)

0.125 2200 0.275

Table 22 Total cost of ammonia in Northwestern Ontario for various scenarios (ammonia production in Alberta)

Cost contribution ($/kg) Hydrocarbon dissociation Steam methane reforming with carbon
in Alberta capture storage (CCS) in Alberta

Production cost 0.3 0.42


Storage cost 0.095 0.095
Transportation cost 0.275 0.275
Total cost of ammonia 0.67 0.79

3.2.6.3.5 Ammonia production in Alberta


If ammonia is produced using hydrocarbon dissociation route in Alberta, the approximate distance of transportation is 2200 km as
shown in Table 21. The total cost of ammonia is calculated and comparatively shown in Table 22 and Fig. 62.
Considering the cost of storage in Alberta and transportation to Northwestern Ontario, the total cost of ammonia increases to
about 0.67 $/kg for hydrocarbon dissociation route as shown in Fig. 62. However, SMR with CCS method represents slightly
higher price corresponding to 0.79 $/kg.
Ammonia Production 89

0.9

0.8

Total cost of ammonia ($/kg)


0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
Hydrocarbon Steam methane Hydrocarbon
dissociation reforming with dissociation in Alberta
in Alberta CCS in Alberta with carbon black sales
Fig. 62 Total cost of ammonia in Northwestern Ontario for different routes (ammonia production in Alberta).

0.8

0.7

0.6
Ammonia cost ($/kg)

0.5

0.4

0.3

0.2

0.1

0
Low-cost Low-cost Wind high Wind
hydropower hydropower pressure electrolysis
high pressure electrolysis electrolysis
electrolysis
Fig. 63 Comparison of ammonia cost produced on-site using wind and hydropower.

Hydrocarbon dissociation also produces carbon black which is a commercial commodity in the market. Per each kilogram of
ammonia produced, about 0.5 kg of carbon black can be obtained from methane dissociation. If the price of carbon black is
assumed to be 1 $/kg, the cost of ammonia for the hydrocarbon dissociation scenario decreases down to 0.17 $/kg.
If the ammonia is produced in a mix pathway such as 50% hydro in Ontario, 25% wind in Newfoundland, 25% hydrocarbon
in Alberta, the total cost of ammonia in a remote community becomes 0.83 $/kg. If the price of carbon black is assumed to be
1 $/kg, the total cost of ammonia in a remote community for the mix scenario decreases down to 0.69 $/kg.

3.2.6.3.6 On-site ammonia production in Northwestern Ontario using renewable resources


As seen in the calculations, the cost of transportation plays a major role in the overall cost of ammonia. Hence, on-site production
of ammonia where it is needed presents promising approach. If it can be realized using low-cost renewable resources depending
on the region, the cost of ammonia can be decreased substantially as illustrated in Fig. 63 and Table 23.

3.2.6.4 Case Study 4 Conclusions


The following concluding remarks can be noted in this analyses and assessment study for alternative ammonia production routes
in Canada.
90 Ammonia Production

Table 23 Comparison of hydrogen and ammonia cost produced on-site using wind and hydropower

Method Electricity price ($/kWh) Hydrogen cost ($/kg) Ammonia cost ($/kg)

Low-cost hydropower high-pressure electrolysis 0.0275 1.73 0.3287


Low-cost hydropower electrolysis 0.0275 2.12 0.424
Wind high-pressure electrolysis 0.063 3.18 0.6042
Wind electrolysis 0.063 3.91 0.782

• The cost of electricity from wind energy corresponds to 4.5 cents/kWh in Newfoundland although lower cost grid electricity
about 1.8 cents/kWh is available for large-scale customers.
• Utilization of high-pressure electrolysis requires less power input resulting in less cost of hydrogen and ammonia.
• The lowest average cost of electricity for large-scale industrial consumers is seen in Winnipeg, Manitoba and Edmonton,
Alberta.
• The current ammonia retail prices continue to decrease. However, ammonia price is strictly dependent on natural gas price
which is not the case for renewable-based options.
• Hydrocarbon dissociation and low-cost hydroelectric routes yield the low ammonia production cost in the range of
0.30–0.40 $/kg.
• There are 58.2 trillion cubic feet of proven reserves in Canada, and there are potentially 113 trillion cubic feet in New-
foundland’s offshore gas fields which can be utilized for the production of ammonia as a sustainable and alternative fuel.
• Renewable-based ammonia production routes have comparative costs with conventional SMR.
• Storage and transportation of ammonia bring additional costs after production which is a significant disadvantage for long
distance transportation.
• If ammonia is produced in Southern Ontario via hydropower and transported to Northwestern Ontario, the total cost of
ammonia is found to be in the range of 0.66–0.75 $/kg.
• When ammonia is produced in Newfoundland via wind energy and transported to Northwestern Ontario, the total cost of
ammonia is found to be in the range of 1.1–1.25 $/kg.
• If ammonia is produced in Alberta via hydrocarbon dissociation and transported to Northwestern Ontario, the total cost of
ammonia is found to be in the range of 0.67–0.79 $/kg. However, since dissociation of methane produced carbon black,
considering the carbon black sales, it decreases down to 0.17–0.29 $/kg.
• Decreasing crude oil prices could be an alternative source of hydrogen and eventually ammonia which are quite satisfactory
energy storage mediums. The cost assessment of hydrocarbon dissociation process requires a more detailed approach because
of the complexity of the method.
• It is seen that on-site production and utilization of ammonia has a significant cost advantage.
• Although hydrocarbon dissociation route is a fossil fuel-based process, the technology is environmentally friendly close to
renewable resources in some environmental impact categories.
• Ammonia production from wind energy is suitable for Newfoundland whereas hydropower in Ontario yields lower product-
ion cost. Additionally, Newfoundland has high potentials of hydropower which can bring even reduced costs compared to wind
energy.
• Considering the recent decreasing electricity prices in Ontario ranging between 1 and 3 cents/kWh, the cost of ammonia
production can decline considerably.
• Production, storage, and transport of ammonia need to be analyzed and assessed at the same time for feasible scenarios.

3.2.7 Future Directions

In this section, a short overview of the current ammonia related projects is given to emphasize the importance of the topic. The
researchers in Colorado School of Mines [48] are in the progress of developing a membrane reactor concept to synthesize
ammonia at ambient pressure. In conventional ammonia production processes, nitrogen and hydrogen compete for identical
catalyst sites, and the attendance of each inhibits the other, with the overall rate reflecting a compromise. The team suggests
decoupling and self-reliantly controlling the nitrogen and hydrogen dissociation by donating one side of the composite membrane
to each. In this way, the catalysts may be independently optimized. Highly active catalysts have been formerly established for
hydrogen dissociation, and the team’s focus is discovering early transition metals which have shown excessive potential as catalysts
for nitrogen dissociation. When achieved, this technology can permit the production of ammonia at ambient pressure, decreasing
the scale and number of steps compulsory in the process [48].
There is a project by University of Glasgow named “first principles design of novel ammonia synthesis catalysts.” This project
applies computational design of metal nitrides to progress novel ammonia synthesis catalysts. When deliberation is given to the
very greatly energy intensive nature of ammonia synthesis as currently experienced on an industrial scale, the financial and
environmental profits to both industry and society as whole resulting from any enhancement is readily superficial. It has been
projected that, worldwide, industrial ammonia synthesis accounts for more than 1% of energy demand. The method to be taken
Ammonia Production 91

involves a combination of computational design and experimental testing and is based upon preceding studies of metal nitride
catalysts which show interesting activity for ammonia synthesis. Metal nitrides possibly comprise “activated” nitrogen within their
structure and it is the reactivity of this lattice nitrogen which could be the key to their high activity [49].
The project lead by the University of Minnesota aims to develop “Small Scale Ammonia Synthesis Using Stranded Wind
Energy.” This project will insert an inorganic absorbent material in the ammonia synthesis loop of a traditional Haber–Bosch
process, which will allow increased single-pass conversion and increase production rates. Importantly, the team hopes that “this
approach will allow for ammonia production at 10 times lower pressure than the Haber–Bosch process, and consequently permit
them to lessen the size of the ammonia plant without losing the efficiencies of scale [50].
RTI International also aims to advance the Haber–Bosch process, with its “innovative renewable energy-based catalytic
ammonia production.” RTI International’s innovation uses a breakthrough catalyst to allow operation at temperatures at least 20%
lower, and with reduced pressures. The benefit of this small-scale reactor is that, different from regular Haber–Bosch, requiring
constant supply of power, the ammonia synthesis process can start and stop in synchronization with intermittent renewable power
sources [50,51].
Research Corporation in West Virginia University [50] is also evolving a chemical process, though less closely related to Haber–
Bosch. The title of the project is “Renewable Energy to Fuels Through Plasma Catalytic Synthesis of Ammonia,” aims to produce
ammonia from hydrogen and nitrogen using a microwave plasma using low temperatures and pressure at five times the con-
version rate of the Haber–Bosch process. Because of the shorter necessary warm-up time, these low temperatures and pressures
make the process amenable to intermittent renewable energy sources [50].
Some researchers in Wichita State University [50] is currently working on “Alkaline Membrane-Based Ammonia Electro-
synthesis With High Efficiency for Renewable and Scalable Liquid Fuel Production.” The researchers at Wichita State University
aim to use a HEM (hydroxide exchange membrane) cell to “increase the selectivity for ammonia product – making it more
efficient – while the device’s tolerance for high electrical current would help lower costs relative to other electrochemical
approaches.” [50].
The researchers in the team of University of Delaware [52] are building a fuel cell with a hydroxide-ion conducting membrane
electrolyte which consumes ammonia directly to generate electricity. Use of such a fuel cell for transportation applications can be
facilitated by a lower cost polymer electrolyte and catalysts capable of oxidizing ammonia effectively around 1001C. The team goal
is an ammonia-fed fuel cell with rapid startup enabled by the low operating temperature [52]. These new projects signify the
importance of novel ammonia production options. As a carbon-free fuel, fertilizer and working fluid, ammonia, can lead to
cleaner communities in the coming future.

3.2.8 Closing Remarks

Ammonia is dominantly produced by SMR in the world. In terms of conventional resources, naphtha, HFO, coal, natural gas, coke
oven gas, and refinery gas can be used as feedstock for ammonia production. It is in the top three chemicals transported annually.
The Haber–Bosch process is the most common method to produce ammonia in the world, however there are other technologies
being developed such as electrochemical ammonia synthesis. This type of ammonia production system uses an electrochemical
process to produce ammonia from nitrogen, water or hydrogen, and electricity. The electrochemical synthesis of NH3 is a
promising alternative to conventional energy intensive NH3 production plants. Using renewable energy resources to drive the
electrochemical NH3 synthesis, the carbon footprint of current NH3 production industry can be significantly lowered. Electro-
chemical NH3 synthesis routes offer higher integrability to stand alone and distributed NH3 production which is a carbon-free fuel
for various sectors. As an alternative and sustainable fuel, ammonia, can be utilized in the diesel generators by minor mod-
ifications. Ammonia has no GHG emission during utilization in the diesel generators, engines and fuel cells. Ammonia can be
produced on-site using renewable energy resources such as wind, solar and hydropower which are already available in many
locations around the world. As the cost of renewable electricity catches the level of conventional electricity, renewable energy-
based ammonia production systems will continue to gain practicality and popularity. As a result, ammonia usage in the com-
munities for power, heating and cooling production will bring significant cost and environmental benefits together with public
satisfaction.

Acknowledgment

The authors acknowledge the support provided by the Natural Sciences and Engineering Research Council of Canada.

References

[1] Industrial Efficiency Technology Database. Ammonia| Measures. Available From: http://ietd.iipnetwork.org/content/ammonia#key-data; 2017 [accessed 07.01.17].
[2] International Energy Agency. Energy technology perspectives pathways to a clean energy system. OECD Publishing; 2012. doi:10.1787/energy_tech-2012-en.
ISBN:9789264174894.
[3] Comparisons. NH3 fuel association. Available From: https://nh3fuelassociation.org/comparisons/; 2017 [accessed 07.01.17].
92 Ammonia Production

[4] Appl M. Ammonia: principles and industrial practice. Weinheim; Chichester: Vch Verlagsgesellschaft Mbh. 1999.
[5] Olson NK, Holbrook J. NH3 – “The other hydrogen”. Ammon Fuel Netw 2007.
[6] Smith AR, Klosek J. A review of air separation technologies and their integration with energy conversion processes. Fuel Process Technol 2001;70(2):115–34.
[7] Bartels JR. A feasibility study of implementing an ammonia economy. Ann Arbor: Iowa State University; 2008.
[8] Rossetti I, Pernicone N, Forni L. Graphitised carbon as support for Ru/C ammonia synthesis catalyst. Catal Today 2005;102–103:219–24.
[9] Jennings JR. Catalytic ammonia synthesis: fundamentals and practice. New York, NY: Springer; 1991.
[10] Erisman JW, Sutton MA, Galloway J, Klimont Z, Winiwarter W. How a century of ammonia synthesis changed the world. Nat Geosci 2008;1(10):636–9.
[11] Marnellos G, Stoukides M. Ammonia synthesis at atmospheric pressure. Science 1998;282(5386):98–100.
[12] Liu R, Xu G. Comparison of electrochemical synthesis of ammonia by using sulfonated polysulfone and Nafion membrane with Sm1.5Sr0.5NiO4. Chin J Chem 2010;28
(2):139–42. WILEY-VCH Verlag.
[13] Xu G, Liu R. Sm1.5Sr0.5MO4 (M ¼ Ni, Co, Fe) cathode catalysts for ammonia synthesis at atmospheric pressure and low temperature. Chin J Chem 2009;27(4):677–80.
WILEY-VCH Verlag.
[14] Ganley JC, Holbrook JH, McKinley DE. Solid state ammonia synthesis. In: Annual NH3 fuel conference, October 15 and 16, 2007, San Francisco, CA; 2007.
[15] Clarke RE, Giddey S, Ciacchi FT. et al. Direct coupling of an electrolyser to a solar PV system for generating hydrogen. Int J Hydrogen Energy 2009;34(6):2531–42.
[16] Xu G, Liu R, Wang J. Electrochemical synthesis of ammonia using a cell with a Nafion membrane and SmFe0.7Cu0.3xNixO3 (x ¼ 0  0.3) cathode at atmospheric
pressure and lower temperature Sci Chin Ser B Chem 2009;52(8):1171–5. SP Science in China Press.
[17] Giddey S, Badwal SPS, Kulkarni A. Review of electrochemical ammonia production technologies and materials. Int J Hydrogen Energy 2013;38(34):14576–94.
[18] Murakami T, Nishikiori T, Nohira T, Ito Y. Electrolytic synthesis of ammonia in molten salts under atmospheric pressure. J Am Chem Soc 2002;125(12):334–5.
[19] Fanning JC. The chemical reduction of nitrate in aqueous solution. Coord Chem Rev 2000;199(1):159–79.
[20] White RE, editor. Modern aspects of electrochemistry New York, NY: Springer; 2009.
[21] Li F-F, Licht S. Advances in understanding the mechanism and improved stability of the synthesis of ammonia from air and water in hydroxide suspensions of nanoscale
Fe2O3 Inorg Chem 2014;53(19):10042–4. American Chemical Society.
[22] Lan R, Irvine JTS, Tao S. Synthesis of ammonia directly from air and water at ambient temperature and pressure Sci Rep 2013;3:1145. Macmillan Publishers Limited.
[23] Licht S, Cui B, Wang B, et al. Ammonia synthesis by N2 and steam electrolysis in molten hydroxide suspensions of nanoscale Fe2O3. Science 2014;345(6197):637–40.
[24] Garagounis I, Kyriakou V, Skodra A, Vasileiou E, Stoukides M. Electrochemical synthesis of ammonia in solid electrolyte cells. Front Energy Res 2014;2:1–10.
[25] Skodra A, Ouzounidou M, Stoukides M. NH3 decomposition in a single-chamber proton conducting cell. Solid State Ion 2006;177(26–32):2217–20.
[26] Serizawa N, Miyashiro H, Takei K. et al. Dissolution behavior of ammonia electrosynthesized in molten LiCl–KCl–CsCl system. J Electrochem Soc 2012;159(4):E87–91.
[27] Pappenfus TM, Lee K, Thoma LM, Dukart CR. Wind to ammonia: electrochemical processes in room temperature ionic liquids ECS Trans 2009 89–93. ECS.
[28] Di J, Chen M, Wang C. et al. Samarium doped ceria–(Li/Na)2CO3 composite electrolyte and its electrochemical properties in low temperature solid oxide fuel cell. J
Power Sources 2010.
[29] Amar IA, Lan R, Petit CTG, Arrighi V, Tao S. Electrochemical synthesis of ammonia based on a carbonate-oxide composite electrolyte. Solid State Ion 2011;182(1):133–8.
[30] Malavasi L, Fisher CAJ, Islam MS. Oxide-ion and proton conducting electrolyte materials for clean energy applications: structural and mechanistic features Chem Soc Rev
2010;39(11):4370–87. The Royal Society of Chemistry.
[31] Tillement O. Solid state ionics electrochemical devices. Solid State Ion 1994;68(1–2):9–33.
[32] Norby T. Solid-state protonic conductors: principles, properties, progress and prospects. Solid State Ion 1999;125(1–4):1–11.
[33] Iwahara H, Yajima T, Hibino T, Ozaki K, Suzuki H. Protonic conduction in calcium, strontium and barium zirconates. Solid State Ion 1993;61(1):65–9.
[34] Nowick AS, Du Y, Liang KC. Some factors that determine proton conductivity in nonstoichiometric complex perovskites. Solid State Ion 1999;125(1–4):303–11.
[35] Badwal SPS, Ciacchi FT. Oxygen-ion conducting electrolyte materials for solid oxide fuel cells Ionics 2000;6(1):1–21. Springer-Verlag.
[36] Grundt T, Christiansen K. Hydrogen by water electrolysis as basis for small scale ammonia production. A comparison with hydrocarbon based technologies Int J
Hydrogen Energy 1982;7(3):247–57. Pergamon.
[37] Michalsky R, Pfromm PH. Chromium as reactant for solar thermochemical synthesis of ammonia from steam, nitrogen, and biomass at atmospheric pressure. Sol Energy
2011;85(11):2642–54.
[38] ISO 14044. Environmental management – Life cycle assessment – Requirements and guidelines. Available From: http://www.iso.org/iso/catalogue_detail?csnumber=38498;
2006 [accessed 07.01.17].
[39] ISO 14040. Environmental management – Life cycle assessment – Principles and framework. Available From: http://www.iso.org/iso/catalogue_detail?csnumber=37456;
2006 [accessed 07.01.17].
[40] Consultants P. SimaPro life cycle analysis version 7.2 (software). PRé sustainability, stationsplein 121, 3818 LE Amersfoort, The Netherlands; 2010.
[41] Barton R. Estimation of costs of heavy vehicle use per vehicle-kilometre final report TP 14556 E; 2006.
[42] DOE Hydrogen and Fuel Cells Program. DOE H2A production analysis. H2A central hydrogen production model version 3.1. Available From: https://www.hydrogen.energy.
gov/h2a_production.html; 2017 [accessed 07.01.17].
[43] Harvey LDD. The potential of wind energy to largely displace existing Canadian fossil fuel and nuclear electricity generation. Energy 2013;50:93–102.
[44] Shankleman J. Green ammonia made with wind is future of fertilizer at Siemens. Bloomberg. Available From: https://www.bloomberg.com/news/articles/2016-04-20/green-
ammonia-made-with-wind-is-future-of-fertilizer-at-siemens; 2017 [accessed 07.01.17].
[45] Market Realist. Ammonia prices fell slightly from the previous week. Available From: http://marketrealist.com/2016/03/weekly-ammonia-price-update-week-ending-march-4-
2016/; 2017 [accessed 07.01.17].
[46] Abánades A, Rathnam RK, Geißler T, et al. Development of methane decarbonisation based on liquid metal technology for CO2-free production of hydrogen. Int J
Hydrogen Energy 2016;41(19):8159–67.
[47] John P. Energy transportation and tanker safety in Canada Fraser Institute, http://www.fraserinstitute.org; 2015.
[48] ARPA. Low cost membrane reactor synthesis of ammonia at moderate conditions, Colorado School of Mines. Available From: https://arpa-e.energy.gov/?q=slick-sheet-
project/ammonia-synthesis-membrane-reactor; 2017 [accessed 14.01.17].
[49] University of Glasgow, EP/L02537X/1. First principles design of novel ammonia synthesis catalysts. Available From: http://gtr.rcuk.ac.uk/; 2017 [accessed 14.01.17].
[50] ARPA-E. Renewable energy to fuels through utilization of energy-dense liquids (REFUEL). Available From: https://arpa-e.energy.gov/; 2017 [accessed 14.01.17].
[51] RTI International. Innovative renewable energy-based catalytic ammonia production 2017;
[52] Fuel Cell Research Lab, Direct ammonia fuel cells for transport applications, Department of Mechanical Engineering, University of Delaware Available From http://www.me.
udel.edu/; 2017 [accessed 14.01.17].

Further Reading
Andersson J, Lundgren J. Techno-economic analysis of ammonia production via integrated biomass gasification. Appl Energy 2014;130:484–90.
Appl M. Ammonia, 3. Production plants. In: Ullmann’s encyclopedia of industrial chemistry. Weinheim, Germany: Wiley-VCH Verlag; 2011. http://dx.doi.org/10.1002/14356007.
o02_o12.
Ammonia Production 93

Appl M. Complete ammonia production plants. In: Ammonia: principles and industrial practice. Wiley-VCH Verlag GmbH; 2007. p.177–204. http://dx.doi.org/10.1002/
9783527613885.ch05.
Appl M. Process steps of ammonia production. In: Ammonia: principles and industrial practice. Wiley-VCH Verlag GmbH; 2007. p.65–176. http://dx.doi.org/10.1002/
9783527613885.ch04.
Bicer Y, Dincer I. Life cycle assessment of nuclear-based hydrogen and ammonia production options: a comparative evaluation Int J Hydrogen Energy 2017 http://dx.
doi.10.1016/j.ijhydene.2017.02.002
Bicer Y, Dincer I, Zamfirescu C, Vezina G, Raso F. Comparative life cycle assessment of various ammonia production methods J Clean Prod 2016;135:1379–95. http://dx.doi.
org/10.1016/j.jclepro.2016.07.023
Chiao L, Rinker RG. A kinetic study of ammonia synthesis: modeling high-pressure steady-state and forced-cycling behavior. Chem Eng Sci 1989;44:9–19.
Dugger GL, Francis EJ, Avery WH. Technical and economic feasibility of Ocean Thermal Energy Conversion. Sol Energy 1978;20:259–74.
Edrisi A, Mansoori Z, Dabir B. Using three chemical looping reactors in ammonia production process – a novel plant configuration for a green production. Int J Hydrogen
Energy 2014;39:8271–82.
Kirova-Yordanova Z. Exergy analysis of industrial ammonia synthesis. Energy 2004;29:2373–84.
Morgan ER, McGowan JG. Techno-economic feasibility study of ammonia plants powered by offshore wind [Ph.D. dissertation]. Amherst: University of Massachusetts; 2013.
Paschkewitz TM. Ammonia production at ambient temperature and pressure: an electrochemical and biological approach [Ph.D. (doctor of Philosophy) thesis]. Iowa: University
of Iowa; 2012.
Shipman MA, Symes MD. Recent progress towards the electrosynthesis of ammonia from sustainable resources Catal Today 2017;286:57–68. http://dx.doi.org/10.1016/j.
cattod.2016.05.008
Siddiq S, Khushnood S, Koreshi ZU, Shah MT, Qureshi AH. Optimal energy recovery from ammonia synthesis in a solar thermal power plant. Arab J Sci Eng
2013;38:2569–77.
Tock L, Maréchal F, Perrenoud M. Thermo-environomic evaluation of the ammonia production. Can J Chem Eng 2015;93:356–62.
van der Ham CJM, Koper MTM, Hetterscheid DGH. Challenges in reduction of dinitrogen by proton and electron transfer. Chem Soc Rev 2014;43:5183–91.
Vitse F, Cooper M, Botte GG. On the use of ammonia electrolysis for hydrogen production. J Power Sources 2005;142:18–26.
Zamfirescu C, Dincer I. Using ammonia as a sustainable fuel. J Power Sources 2008;185:459–65.

Relevant Websites
https://arpa-e.energy.gov/?q=now-leaving
Advanced Research Projects Agency–Energy (ARPA-E).
http://www.ammoniaenergy.org
Ammonia Energy.
https://ammoniaindustry.com/
Ammonia Industry.
https://www.bre.com/Ammonia-Production.aspx
Bryan Research & Engineering, Inc.
https://www.ohio.edu/engineering/ceer/ammonia-electrolysis.cfm
Center for Electrochemical Engineering Research (CEER) – Ohio University.
https://www.ceramatec.com/
Ceramatec, Inc.
https://chemengineering.wikispaces.com/Ammonia+production
ChemEngineering.
https://cleantechnica.com/2016/05/09/researchers-develop-method-producing-ammonia-sunlight/
CleanTechnica.
http://www.csiro.au/
Commonwealth Scientific and Industrial Research Organisation (CSIRO).
https://www.canada.ca/en/health-canada/services/publications/healthy-living/guidelines-canadian-drinking-water-quality-guideline-technical-document-ammonia.html
Government of Canada.
https://www.topsoe.com/forums-research/research-activities-papers/ammonia
Haldor Topsoe.
http://www.nh3fuel.com/
Hydrofuel Inc.
https://www.ipni.net/
International Plant Nutrition Institute (IPNI).
http://www.che.ksu.edu/research/solar_processing/
Kansas State University.
https://www.netl.doe.gov/research/coal/energy-systems/gasification/gasifipedia/fertilizer-commercial-technologies
National Energy Technology Laboratory (NETL).
http://www.nrel.gov/news/press/2016/25674
National Renewable Energy Laboratory (NREL).
https://www.health.ny.gov/environmental/emergency/chemical_terrorism/ammonia_tech.htm
New York State.
http://www.nh3car.com/
NH3CAR.com.
https://www.nh3fuelassociation.org/
NH3 Fuel Association.
https://www.pnnl.gov/science/highlights/highlight.asp?id=3951
Pacific Northwest National Laboratory (PNNL).
https://www.worldfertilizer.com/
Palladian Publications Ltd.
http://www.potashcorp.com/
PotashCorp.
94 Ammonia Production

http://www.spg-corp.com/clean-energy-power-generation.html
Space Propulsion Group (SPG).
https://www.thyssenkrupp-industrial-solutions.com/en/products-and-services/fertilizer-plants/ammonia-plants-by-uhde/ammonia-plants-500mtpd/the-uhde-ammonia-processes/
Thyssenkrupp AG.
http://www.titech.ac.jp/english/about/stories/ammonia_synthesis.html
Tokyo Institute of Technology.
http://www.uni-regensburg.de/chemistry-pharmacy/inorganic-chemistry-korber/index.html
Universität Regensburg.
https://wcroc.cfans.umn.edu/displacing-diesel-fuel
West Central Research and Outreach Center – University of Minnesota.
https://wcroc.cfans.umn.edu/research-programs/renewable-energy/wind-hydrogen
West Central Research and Outreach Center – University of Minnesota.
http://www.yara.com/
Yara.

You might also like