Sie sind auf Seite 1von 39

2.

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

2.1.1 Introduction 2
2.1.1.1 Ammonia Production and Transport 5
2.1.1.2 Ammonia Storage 6
2.1.1.3 Ammonia Utilization 8
2.1.1.3.1 Ammonia in Heating, Ventilation and Air Conditioning applications 11
2.1.1.3.2 Ammonia as both fuel and refrigerant 12
2.1.1.3.2.1 Thermo-catalytic decomposition of ammonia 13
2.1.1.3.3 Ammonia and urea 14
2.1.2 Analysis and Assessment 14
2.1.2.1 Ammonia in Direct Ammonia Solid Oxide Fuel Cell 14
2.1.2.2 Urea in Direct Urea Solid Oxide Fuel Cell 15
2.1.2.3 Utilization of Fuel and Oxidant 15
2.1.2.4 Electrochemical Analysis 15
2.1.2.4.1 Activation overpotential 15
2.1.2.4.2 Ohmic overpotential 16
2.1.2.4.3 Concentration overpotential 16
2.1.2.4.4 Electrochemical efficiency 17
2.1.2.5 Illustrative Example 17
2.1.3 Case Study 1: Ammonia Utilization in Road Vehicles 20
2.1.3.1 Materials and Methods 20
2.1.3.2 Systems Description 22
2.1.3.2.1 Passenger car manufacturing 22
2.1.3.2.2 Maintenance 23
2.1.3.2.3 Disposal of the vehicles 23
2.1.3.2.4 Operation of vehicles 23
2.1.3.3 Case Study 1 Results and Discussion 24
2.1.3.4 Case Study 1 Conclusions 28
2.1.4 Case Study 2: Ammonia Utilization in Maritime Applications 29
2.1.4.1 Introduction 29
2.1.4.2 Pathways for Production of Ammonia 29
2.1.4.3 Material and Methods 30
2.1.4.4 Life Cycle Assessment Phases 31
2.1.4.4.1 Transoceanic freight ship manufacture 31
2.1.4.4.2 Maintenance of transoceanic freight ship 31
2.1.4.4.3 Port facilities 31
2.1.4.4.4 Maintenance and operation of the port 31
2.1.4.4.5 Operation of the transoceanic freight ship and transport for 1 tkm 32
2.1.4.5 Case Study 2 Results and Discussion 32
2.1.4.6 Case Study 2 Conclusions 35
2.1.5 Closing Remarks 36
Acknowledgment 37
References 37
Further Reading 38
Relevant Websites 39

Nomenclature E Open-circuit voltage (V)


Bo Permeability of the porous electrode F Faraday constant
Deff
i;j Effective binary diffusion j Current density
Deff
i;k Effective Knudsen diffusion N Molar flux
Uf Utilization of fuel P Pressure (bar)
Uo Utilization of air R Gas constant
ΔH Enthalpy change (kJ kg1) T Temperature (K)

Comprehensive Energy Systems, Volume 2 doi:10.1016/B978-0-12-809597-3.00201-7 1


2 Ammonia

V Working potential or voltage (V) z Electron number


y Molar fraction of chemical species

Acronyms LPG Liquefied petroleum gas


ADF Abiotic depletion factor MSWI Municipal waste incineration plant
CFC Chlorofluorocarbon MTPM Mean transport pore model
CML Center of Environmental Science of Leiden Ni-YSZ Nickel-yittria stabilized zirconia
University NM Non-methane
DA Direct ammonia NMI Nautical mile
DB Dichlorobenzene OCV Open cell voltage
DeNOx Destruction of nitrous oxides OEM Original equipment manufacturer
DGM Dusty-gas model PAH Polycyclic aromatic hydrocarbon
DU Direct urea PTH Pump to haul
DWT Dead weight tonnage PV Photovoltaic
EV Electric vehicle RCV Refuse collection vehicles
FFE Feed plus fuel energy RSZ Reduced-speed zones
GEM Gibbs energy minimization SMR Steam methane reforming
GHG Greenhouse gas SOFC Solid oxide fuel cell
GREET Greenhouse gases, regulated emissions, and SOFC-H Solid oxide fuel cell with hydrogen proton
energy use in transportation (H þ ) conducting electrolyte
GWP Global warming potential SOFC-O Solid oxide fuel cell with oxygen ion (O2)
HEV Hybrid electric vehicle conducting electrolyte
HFC Hydrofluorocarbon STP Standard temperature of 298 K and pressure of
HHV Higher heating value 101.325 Kpa
HVAC Heating, ventilation and air conditioning SUR Steam-to-urea ratio
ICE Internal combustion engines TKM Tonne-kilometer
IPCC Intergovernmental panel on climate change TPB Triple phase boundary
ISO International organization for standards VOC Volatile organic compounds
KN Knot WTH Well to haul
LCA Life cycle assessment WTP Wheel to pump
LHV Lower heating value YSZ Yittria-stabilized zirconia

Greek Letters t Tortuosity of electrode


a Charge transfer coefficient jact,an Anode activation overpotential (V)
G Dimensionless temperature jact,ca Cathode activation overpotential (V)
δ Electrolyte thickness (M) jconc,an Anode concentration overpotential (V)
D Net change of quantity jconc,ca Cathode concentration overpotential (V)
e Porosity of electrode jO Ohmic overpotential of electrolyte (V)
e/κ Lennard-Jones temperature parameter (K) OD,ij Collision integral
r Density (kg m3) Z Efficiency
sij Mean characteristic length m Viscosity

Subscripts and superscripts f Fuel


Act Activation i Chemical species
An Anode Mix Mixture
Ca Cathode o Oxidant
Conc Concentration r Reactants
Elc Electrochemical rev Reversible

2.1.1 Introduction

Inevitable issues of fossil fuels and their extensive use in many sectors, including transportation domain, have resulted in
damaging effects on human health and well being as well as the environment. Thus, there is an urgent need to develop and
Ammonia 3

implement some efficient, effective and environmentally benign potential replacements. The relevant studies suggest switching
from the fossil fuels to renewable energy sources and carbon-free fuels, such as hydrogen and ammonia [1]. In this respect, there is
a critical need for main investments for the growth of competitive hydrogen manufacture, delivery, and the particularly storing
technologies that are not commercially fully viable and well developed. There is a need to further study and develop viable
technologies for mass implementations [2]. Therefore, there is a drawback regarding hydrogen that hydrogen delivers very little
quantities of energy per unit volume as compared to the conventional fuels used in various transportation fuels. On top of this, the
growth of hydrogen supply substructure (essentially the infrastructure required) will require some specific safety measures
undertaken particularly on its flammability and explosivity risks since hydrogen is recognized as volatile energy carrier with the
invisible flame [3].
Ammonia is expected to overcome the above mentioned hydrogen related challenges by offering a potential solution to reduce
these challenges, and projected to be a potential hydrogen carrier with high hydrogen content in the near future. In recent years,
expectations are rising for hydrogen and hydrogen carriers as a medium for storage and transportation of energy in the mass
introduction and use of renewables. Both storage and transport of hydrogen are considered an important issue since hydrogen is in
gas form under normal temperature and pressure. Hydrogen carriers are mediums that convert hydrogen into chemical substances
containing large amounts of hydrogen, to simplify storage and transport processes. Hydrogen carriers include ammonia synthe-
sized from nitrogen and hydrogen that can be used for direct combustion, cooling, fuel cells, power generation, etc. Ammonia
becomes a primary hydrogen carrier that does not contain any carbon atoms and has a high hydrogen ratio. Therefore, it is
evaluated as a clean power-generating fuel. Since ammonia produces mainly water and nitrogen on combustion, replacing a part
of conventional fuel with ammonia will have a large effect in reducing carbon dioxide emissions.
In this regard, a unique list of the specific features of ammonia (NH3) is as follows:

• It consists of one nitrogen atom from air separation and three hydrogen atoms from any conventional or renewable resources.
• It is the second largest synthesized industrial chemical in the world.
• It is a significant hydrogen carrier and transportation fuel that does not contain any carbon atoms and has a high hydrogen
ratio.
• It does not emit direct greenhouse gas emission during its deployment.
• It can be used as solid and/or liquid for many purposes.
• It can be stored and transported under relatively lower pressures compared to hydrogen.
• It can be produced from a various type of resources ranging from oil sands to renewables.
• It is a suitable fuel to be transferred using steel pipelines with minor modifications which are currently used for natural gas and
oil.
• It can be used in all types of combustion engines, gas turbines, burners as a sustainable fuel with only small modifications and
directly in fuel cells which is a critical advantage compared to other type of fuels.
• It brings a non-centralized power generation via fuel cells, stationary generators, furnaces/boilers, and enables smart grid
applications.
• It can be used as a refrigerant for cooling in the vehicles and other applications.

Ammonia can be produced by extracting nitrogen from air and hydrogen from water and combining them with the presence of
any energy source, including, especially, the renewables. Presently, more than 90% of the ammonia manufacture in the world is
done by the Haber–Bosch synthesis process as it was developed in 1913 [4,5]. The Haber–Bosch process combines hydrogen and
nitrogen together with an iron-oxide catalyst under very high pressure and elevated temperatures. On the other hand, new expertise
such as solid state synthesis and electrochemical processes are currently being developed to reduce the cost and to improve the
efficiency of ammonia production processes.
Ammonia is one of the major synthesized industrial chemicals on Earth. Ammonia production consumes nearly 1.2% of total
primary energy and subsidizes about 1% of GHGs emissions [6]. About 1.5 t of CO2 is released into the environment during the
production of 1 t of ammonia depending on the location of the plant [7]. Regarding conventional sources, naphtha, heavy fuel oil,
coal, natural gas coke oven gas, and refinery gas can be used as feedstock for ammonia manufacture. Natural gas is the main
feedstock used for making ammonia in Canada, and worldwide. Eleven ammonia plants are operating in Canada, producing a
yearly average of 4–5 million metric tons per plant [8]. The United States is accepted as the biggest ammonia importing country
which corresponds to about 35–40% of world trade. On the other hand, Europe accounts for approximately 25% of commerce
while they produce ammonia with a higher cost. In the coming future, it is expected that the majority of import development will
be in Asia because of industrial uses and fertilizer products.
Though there are a few practical techniques for ammonia synthesis, commonly there is one ammonia synthesis technique
available in the world named as the Haber–Bosch process [9] whereas the electrochemical ammonia synthesis processes are also
under investigation. In both approaches, nitrogen is supplied through the air separation process. Cryogenic air separation is
presently the most efficient and cost-effective technology for generating a large amount of oxygen, nitrogen, and argon [10]. Using
cryogenic technology, nitrogen can also be produced in high purity which can further be utilized as a useful by-product stream at
quite small incremental price. Among other air separation processes, cryogenic air separation has most mature and developed
technology. Since ammonia is produced in high amounts, required nitrogen should be generated in a low cost and high efficient.
The required electricity could be supplied either from conventional or alternative sources. The Haber–Bosch process is an
4 Ammonia

Table 1 Thermo-physical properties of ammonia

Property Liquid

Color Colorless
Density (01C, 101.3 kPa) 638.6 kg m3
Density (  33.431C, 101.3 kPa) 682 kg m3
Boiling point (101.3 kPa)  33.431C
Melting point  77.711C
Critical temperature 132.41C
Critical pressure 11.28 MPa.
Critical viscosity 23.90  103 mPa s
Lower heating value, LHV (MJ kg1) 18.57
Higher heating value, HHV (MJ kg1) 22.54

Source: Data from Kaudy L, Rounsaville JF, Schulz G. Ullmann’s encyclopedia of industrial
chemistry. Weinheim: VCH; 1987.

Table 2 Some properties of ammonia, hydrogen and other conventional fuels

Feature Gasoline Diesel Natural gas H2 NH3

Flammability limit, volumes (% in air) 1.4–7.6 0.6–5.5 5–15 4–75 16–25


Auto-ignition temperature (1C) 300 230 450 571 651
Peak flame temperature (1C) 1977 2053 1884 2000 1850

Source: Data from Zamfirescu C, Dincer I. Ammonia as a green fuel and hydrogen source for vehicular applications. Fuel Process Technol
2009;90:729–37.

exothermic process that combines hydrogen and nitrogen in 3:1 ratio to produce ammonia. The reaction is facilitated by catalyst
and the optimal temperature range is 450–6001C [11,12].
Ammonia (NH3), which has a great content of three atoms of hydrogen per unit of capacity, has occasionally been used in the
history as a fuel for internal combustion engines (ICEs) and fuel cells [13]. One mole of ammonia comprises 1.5 mol of hydrogen
which is 17.8% by weight or 108 kg H2 m3 entrenched in liquid ammonia at 201C and 8.6 bar. It is remarkable to know that this
density is four times greater than that of the most advanced storing approaches in metal hydrides which extend about 25 kg H2
m3 [14]. The physical properties of ammonia are listed in Table 1.
NH3 is one of the most attractive fuels due to the following features:

• It has a high octane rate of 110–130 [15] and therefore is a decent fuel for ICEs.
• It can be thermally decomposed into hydrogen and nitrogen using low energy, i.e., approximately 12% from the higher heating
value (HHV) [2] to generate hydrogen for fuel cells.
• The delivery substructure already exists for ammonia [16] to deliver it in quantities superior than 100 million tons annually
[17] or more.
• It is still safer than many fuels and hydrogen due to the following possessions:
✓ If escapes into the air, it disperses fast because its density is lighter than that of air.
✓ It is self-alarming: nose can notice any leak in concentrations as low as 5 ppm [9].
✓ It has a slight flammability range and so, it is typically evaluated nonflammable and giving no explosion danger when
correctly conveyed; this detail is evident from the data listed in Table 2 which were compiled from various sources [18–20].
Although there are some flammability and toxicity concerns linked to NH3 and its use, these challenges are easy to overcome
with the appropriate safety measures and control. That is why this has been the most widely used refrigerant and working fluid
since the industrial revolution. Especially, there are some side effects with every refrigerant, working fluid and fuel. Nonetheless,
ammonia appears to be one of the most environmentally friendly refrigerant and working fluid and hence fuel while con-
sidering the ammonia produced by renewable hydrogen and nitrogen (from the air). If we further elaborate, such difficulties
have mainly been addressed, and these apparent disadvantages are compensated by the well-established knowledge in NH3
treatment, storing and use in numerous forms (i.e., gaseous, liquid as well as solid), particularly in agriculture, chemical, and
cooling applications.
Some quantities of NOx will consequence due to the extra nitrogen presence in the burning cavity if NH3 is used as fuel for
ICEs. It is also stimulating to note that NH3 is a reduction agent for the NOx, typically in combustion releases. The reaction of NOx
with ammonia over catalysts produces only steam and nitrogen. An average car needs approximately 30 mL of NH3 per 100 km to
neutralize any NOx emissions [21]. If the vehicles run on NH3 as fuel, this amount is unimportant with respect to the fuel tank
volume.
Ammonia 5

Table 3 Comparison of ammonia with other fuels

Fuel/storage P (bar) r Density HHV HHV per Energy per Cost per Cost per volume Cost per
(kg m3) (MJ kg1) volume volume mass ($ m3) energy
(GJ m3) (GJ m3) ($ kg1) ($ GJ1)

Gasoline, C8H18/liquid 1 736 46.7 34.4 34.4 1.03 754.86 21.97


CNG, CH4/integrated storage 250 188 42.5 10.4 7.8 0.91 170.60 21.29
LPG, C3H8/pressurized tank 14 388 48.9 19 11.7 1.06 413.66 21.74
Methanol, CH3OH/liquid 1 786 14.3 11.2 9.6 0.41 317.80 28.31
Hydrogen, H2/metal hydrides 14 25 142 3.6 3 3.02 75.49 21.29
Ammonia, NH3/pressurized tank 10 603 22.5 13.6 11.9 0.23 136.63 10.04
Ammonia, NH3/metal amines 1 610 17.1 10.4 8.5 0.23 138.14 13.21

Source: Reproduced from Zamfirescu C, Dincer I. Ammonia as a green fuel and hydrogen source for vehicular applications. Fuel Process Technol 2009;90:729–37.

Ammonia is considered as a possible working fluid for thermodynamic cycles, refrigeration, heating, and power [22,23]. It is
used for exhaust gases to drive automotive absorption refrigeration system [24]. More lately, Zamfirescu and Dincer [25–27] have
examined the option of using the on-board ammonia as refrigerant while it is spent as fuel for vehicle propulsion. Ammonia also
seems to be a medium for thermochemical storing of energy as confirmed by numerous works [28,29]. Consequently, ammonia
can play a critical part in vehicular applications, as fuel, a hydrogen source, working fluid/refrigerant, NOx reduction agent, or
perhaps as an energy storage intermediate.
In Table 3, it is listed the fuel and the type of storing option, the fuel pressure in the storing tank, and the fuel density in the full
tank. The exact fuel energy per mass is given regarding fuel’s HHV. The volumetric energy of the fuel is found by multiplying the HHV
with the density value listed in the third column. Ammonia’s HHV is around half of the one of gasoline, and its density is also
inferior. Therefore, liquid NH3 stores 2.5 fewer energy per unit capacity than gasoline. If the NH3 is stored in the form of hex-
aamminemagnesium chloride [30] to remove the hazard related to its toxicity, the energetic cost to pay for discharging ammonia
reduces its HHV. The manufacture of ammonia from fossil fuels is in numerous features similar to hydrogen manufacture since it
includes gasification to produce syngas, gas cleaning and CO2 elimination, compression of the reactants and catalytic conversion, and
ammonia separation through condensation. An extremely energy consuming constituent of the ammonia manufacture is considered
as the makeup gas compression which is needed to ease the synthesis. This apparent disadvantage is compensated by an efficient
synthesis procedure that is likely at high pressure. The procedure is eased even more by the facile separation of NH3 product that is
made at no extra costs by ammonia condensation and heat recovery. Thus ammonia is continuously drained from the synthesis loop
such that the chemical equilibrium in the reactor is shifted toward ammonia formation. Furthermore, ammonia synthesis is an
exothermic process, and contemporary technologies use heat recovery to reduce the production prices [31].
Strickland [32] pointed even since 1980 that amongst other fuels, counting liquid hydrogen and methanol, ammonia (NH3) has
some excellent advantages which make it the neighboring one to gasoline. Numerous other new studies have proposed NH3 as a
possible hydrogen storing medium [9,16,33]. A significant advance in ammonia synthesis was noticed in 1931 when Shell Chemicals
produced NH3 from natural gas for the first time [27]. Starting with the natural gas crisis in 1970 ammonia manufacturers became
worried to advance and develop methods to efficiently produce ammonia from other fossil fuels, counting coal [34]. In order to
decompose ammonia and separate the hydrogen from nitrogen at the same time, Skodras et al. [35] industrialized a catalytic
membrane decomposition unit: after the decomposition procedure which take place typically at the catalytic membrane surface, only
hydrogen permits through the membrane; thus small traces of unreacted ammonia and the produced nitrogen are separated from the
pure H2 stream. As an alternative option, hydrogen can also be obtained via ammonia electrolysis [36]. This will definitely make it
more pure, more cost effective, more commercially viable and reliable, and more environmentally benign than the hydrogen
produced through some conventional techniques such as hydrogen production from natural gas.

2.1.1.1 Ammonia Production and Transport


The annual increase in ammonia production covering the period from 2002 to 2015 is shown in Fig. 1. Regarding conventional
resources, naphtha, heavy fuel oil, coal, natural gas coke oven gas and refinery gas can be used as feedstock for ammonia production.
Natural gas is the primary feedstock used for producing ammonia in worldwide as shown in Fig. 2. Ammonia production plants in
Canada are ranked internationally as having the highest feed plus fuel energy (FFE) plant efficiency which consumes a typical 33.8 GJ
natural gas per ton of produced ammonia [37]. In comparison, the world FFE average is of 38.6 GJ t1 NH3. FFE is related to the CO2
generated by the ammonia plant, whereas net energy efficiencies contain electrical consumption and modifications for other energy
debits and credits, which may have connected offsite CO2 emissions not directly from the ammonia plant [37]. In China, coal is
intensively used and is generally characterized by high energy requirements [38]. In 2010, ammonia plants in China had a projected
average energy intensity of 49.1 GJ t1 NH3. About 75% of the ammonia in China was produced with the coal-based method. The
average energy intensity of the Chinese coal-based ammonia-producing plants is 54 GJ t1 NH3 [38]. Natural gas costs constitute
70–90% of the production cost of ammonia. Since ammonia production is based on natural gas in the steam methane reforming
(SMR) method, if natural gas prices rise, production costs for ammonia increase in parallel. For the Haber–Bosch process, production
of ammonia is based on various hydrogen production techniques as shown in Fig. 3.
6 Ammonia

200
180

Global ammonia production


160
140

(million tonnes)
120
100
80
60
40
20
0
2002 2004 2006 2008 2010 2012 2014 2016
Year

Fig. 1 World ammonia production growth. Data from Yara Fertilizer Industry Handbook. Available From: http://www.yara.com; 2014.

Naphta
Others
Fuel oil 1%
1%
4%

Coal
22%

Natural gas
72%

Fig. 2 Sources of global ammonia production based on feedstock use. Data from International Energy Agency. Energy Technology Perspectives
2012; Pathways to a Clean Energy System, France. ISBN: 978-92-64-17488-7.

The storage and distribution structure of ammonia are comparable to the liquefied petroleum gas (LPG) process. Under
medium pressures (in the range of 5–15 bar), both of the materials are in liquid state. This qualifies the significant advantage
because of storing opportunities. Nowadays, vehicles running on propane are commonly recognized and used by the public since
their on-board storage is possible and it is a good case for ammonia-fueled vehicle opportunities since the storage and risk
characteristics of both substances are similar to each other. An ammonia pipeline from the Gulf of Mexico to Minnesota and with
divisions to Ohio and Texas has served the ammonia industry for many years. It indicates that there is a working ammonia
pipeline transportation which can be spread overall the world. The potential of using ammonia in many applications will be
dependent on the accessibility of ammonia in the cities. Ammonia is a suitable material to be transported using steel pipelines
with slight adjustments which are presently used for natural gas and oil. In this way, the problem of accessibility of ammonia will
be abolished. A pipeline can deliver almost 50% more energy when conveying liquid ammonia than carrying compressed natural
gas because of the volumetric energy densities [39].

2.1.1.2 Ammonia Storage


Ammonia can be stored in two different ways, pressurized or at low temperature. Pressurized storage keeps ammonia in a liquid
phase having a pressure above 8.6 bar at ambient temperature (201C), but ammonia is usually stored at 17 bar to keep ammonia
in liquid phase if ambient temperature increases. The energy density of the liquid ammonia stored pressurized is 13.77 MJ L1. A
rule of thumb is that 2.8 t of ammonia may be retained per ton steel. This storage does not require energy to maintain the
pressurized state. Low-temperature storage is usually employed for large-scale applications.
This type of storage requires energy to maintain its low temperature and thereby avoid boil-off due to ambient temperature.
Lower capital cost is the reason why low temperature is preferred for large scale storage. The energy density of the liquid ammonia
Ammonia 7

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)

Internal combustion
engine Fuel cell systems

Fig. 3 Ammonia production and usage routes.

10,000
10,000
Electricity: batteries Hydrogen (70 bar) ICEHV

8000 Hydrogen (70 bar) FCHEV CNG

Ammonia (20 bar) Gasoline, diesel


6000
Cost ($)

4000
4000
3000

2000

300 300 100


0
Fig. 4 Estimated OEM costs of on-board ammonia storage tanks for one personal vehicle with 482 km range. Data from Leighty Bill. Energy
storage with anhydrous ammonia: comparison with other energy storage. In: 2008 annual NH3 fuel conference NH3 fuel association, Minneapolis;
2008.

stored in this way is 15.37 MJ L1 compared to 13.77 MJ L1 for pressurized storage. If storage time is assumed 182 days
representing a period between winter and summer, will give a storage cost of 4.03 $ GJ1 for ammonia. It can be mentioned that
this cost is much lower compared to hydrogen storage that costs 98.74 $ GJ1. As illustrated in Fig. 4 the estimated cost of a storage
tank on a personal vehicle is lowest for ammonia after standard and conventional gasoline/diesel tanks for a 482 km range. The
hydrogen density is highest for ammonia among other fuels as illustrated in Fig. 5.
Because the main application of ammonia is fertilizer, large-capacity seasonal storing tanks are already developed. Ammonia
demands peak throughout the summer when it must be spread on farming fields. Ammonia is produced throughout the year, and
the winter’s manufacture is stored for the summer period. Tanks with a volume of 15,000–60,000 m3 were constructed before the
1970s [40]. Ammonia is kept in the chilled state at ambient pressure and at its standard boiling point, which is  331C. The tanks
are cylindrical with a 38–52 m inner diameter and 18–32 m of height. In order to compensate for the heat diffusions, the whole
building is well insulated (a double-wall technology is used) and compressors are employed to eliminate the heat by simulating a
cooling plant for which the tank plays the role of an evaporator. Essentially, ammonia vapors existing above the liquid are aspired
8 Ammonia

140 136
126 Hydrogen density LHV 120.1
120 116
110
106

Lower heating value (MJ kg–1)


Hydrogen density (kg H2 m–3)
102 98
100

80 70

60
50.1 45.8
42.8 42.5
40
27
18.6 20.1
20

0
Ammonia Diesel Methane Gasoline Propane Ethanol Methanol Hydrogen

Hydrogen density 136 126 116 110 106 102 98 70


LHV 18.6 42.8 50.1 42.5 45.8 27 20.1 120.1

Fig. 5 Comparison of hydrogen density and lower heating value (LHV) values of various fuels.

by the compressors and delivered at high pressure where the vapors are condensed, and the liquid is returned to the tank. In this
way, the temperature and the pressure in the tank are kept constant [40,41].
Ammonia can be stored onboard of a car in pressurized cylinders in an anhydrous form or in some chemical forms such as
metal amines or ammonia boranes, which are produced by means of lately developed physical–chemical reversible methods
[30,41]. In this technology, ammonia is adsorbed on a porous metal–amine complex, for example, hexaaminemagnesium
chloride, Mg(NH3)6Cl2; to do this, NH3 is passed over an anhydrous magnesium chloride (MgCl2) powder at room temperature.
The absorption and desorption of ammonia in and from MgCl2 are entirely reversible. The metal amine can be formed in the
desired form and can stock 0.09 kgH2 kg1 and 100 kgH2 m3.

2.1.1.3 Ammonia Utilization


The ability to use one fuel in all types of combustion engines, gas turbines, burners, and directly in fuel cells is a tremendous
advantage. Storage and delivery infrastructure would be significantly reduced if ammonia is employed rather than hydrogen. NH3
is one of a very short list of fuels that can be used in nearly every type of engine and gas burner with only minor modifications. Gas
burners can be equipped with in-line partial reformers to split approximately 5% of the NH3 into hydrogen. This mixture produces
a robust, unpolluted burning open flame. One pipeline to a home could provide NH3 to furnaces/boilers, fuel cells, stationary
generators and even vehicles. Due to the very minor enthalpy of reforming exhibited by NH3, it can easily be reformed to hydrogen
for any application that would require hydrogen. Relatively minor modifications allow efficient use of ammonia as a fuel in diesel
engines; high compression ratio spark ignition engines can produce astounding efficiencies of over 50% using NH3 fuel; direct
ammonia fuel cells promise to be low-cost, robust, and very efficient; NH3 is also a very suitable fuel for use in solid oxide fuel cell
(SOFC) and gas turbines. These medium-temperature (approximately 4001C) fuel cells promise to be low-cost, highly efficient,
and very robust [39].
The global ammonia demand is forecasted to grow at an average annual rate of approximately 3% over the next 5 years. The
historical growth rate was 1%. Therefore, currently, it is 2% above. The global ammonia consumption amounts are shown in
Fig. 6. Solid agricultural materials are expected to drive this growth as fertilizer uses account for approximately 80% of global
ammonia demand [42]. The global ammonia exports and imports based on the selected years are illustrated in Fig. 7. The United
States and Asia have quite close import rates whereas Western Europe’s imports are almost half of Asia. In recent years, the export
of ammonia from Africa and Middle East increases gradually. The United States is recognized as the largest ammonia importer and
typically accounts for approximately 35–40% of world trade. Europe, a higher-cost producer, accounts for roughly 25% of
commerce. The majority of growth in imports is expected in Asian countries, for industrial uses and the production of fertilizer
products.
The physical properties of ammonia require high-pressure containers, making it a little costly and difficult to transport. Most of
the ammonia is consumed close to where it is produced as illustated in Fig. 8. The domestic sales represent approximately 88% of
world ammonia trade. Asia is the main ammonia trading country more than sum of other continents. Almost 53% of ammonia is
currently used as fertilizer in the United States as shown in Fig. 9. The direct applications constitute only quarter of whole usage.
Ammonia 9

Million tonnes Urea Other fertilizer DAP/MAP Industrial


200
180
160
140
120
100
80
60
40
20
0
2001 2003 2005 2007 2009 2011 2013 2015
Fig. 6 World ammonia consumption and distribution. Data from PotashCorp Integrated Annual Report. Annual integrated report. Available From:
http://www.potashcorp.com/irc/nitrogen; 2015 [accessed 07.01.17].

Imports Exports
30 30 Trinidad FSU
US W. Europe
Asia Other Africa Others
25 25 Middle east

20 20
Million tonnes

Million tonnes

15 15

10 10

5 5

0 0
2005 2007 2009 2011 2013 2015 2005 2007 2009 2011 2013 2015
Fig. 7 World ammonia trade shares. Data from PotashCorp Integrated Annual Report. Annual integrated report. Available From: http://www.
potashcorp.com/irc/nitrogen; 2015 [accessed 07.01.17].

However, when the world ammonia usage is considered, direct applications represent only 4% of overall ammonia consumption
as illustrated in Fig. 10. The lack of ammonia using devices and equipment leads indirect applications.
Utilization of ammonia in household applications is also possible in various ways. In an ammonia economy, the readiness of a
pipeline to the residential area could source ammonia to fuel cells, stationary generators, furnaces/boilers, and even vehicles which
will bring a non-centralized power production and allow smart grid applications [39]. Decentralized power generation and
utilization is one of the solutions for transmission lines. Ammonia can play a crucial role in this process since it has multiple usage
options [43].
It is emphasized that the physical characteristics of ammonia are close to propane. The capability to convert a liquid at
adequate pressure permits ammonia to store more hydrogen per unit volume than compressed hydrogen/cryogenic liquid
hydrogen. Besides having significant advantages in storing and transporting hydrogen, ammonia may also be burned directly in
ICE. Compared to gasoline vehicles, ammonia-fueled vehicles do not produce direct CO2 emission during operation. However, it
is important to determine not only direct emissions associated with vehicle operation but also total energy cycle emissions related
to fueling the vehicles. Furthermore, ammonia can be produced at locations where oil and natural gas extraction wells are located.
In this way, generated CO2 can be reinjected into the ground for sequestration or can be reacted with ammonia for urea
production. Ammonia can then be easily transferred through pipelines, railway cars, and ships by delivering to consumption area
where it may be utilized as a source of hydrogen, chemical substance, and fertilizer for agriculture, fuel for transportation and
power generation sector, working fluid or refrigerant. Ammonia can be utilized in many transportation applications as shown in
Fig. 11.
10 Ammonia

Production - Million tonnes 2010 2015F

Asia

FSU
Exports
Middle East
12%
Europe

North America

Latin America
88%
Africa
Domestic sales
Oceania

0 10 20 30 40 50 60 70 80 90
Fig. 8 Domestic- and export-based world ammonia production profile. Data from PotashCorp Integrated Annual Report. Annual integrated report.
Available From: http://www.potashcorp.com/irc/nitrogen; 2015 [accessed 07.01.17].

Non-fertilizer

23%

53% Upgraded
fertilizers

24%
Direct
application

Fig. 9 Ammonia consumption in the United States for industrial and fertilizer purposes. Data from PotashCorp Integrated Annual Report. Annual
integrated report. Available From: http://www.potashcorp.com/irc/nitrogen; 2015 [accessed 07.01.17].

Non fertilizer
19%

Direct application
4%
Urea
48%
DAP/MAP
7%

Other fertilizer
14%

Ammonium nitrate
8%
Fig. 10 World ammonia usage, average of 2010–2013. Data from PotashCorp Integrated Annual Report. Annual integrated report. Available From:
http://www.potashcorp.com/irc/nitrogen; 2015 [accessed 07.01.17].
Ammonia 11

Ammonia
synthesis

Ammonia
storage

Compression
Spark ignition
Fuel cells ignition
engines
engines

Ammonia dual Ammonia duel


Ammonia Ammonia
fuel fuel

Ammonia Ammonia Ammonia


Ammonia and Ammonia and
and and and
gasoline hydrogen
diesel DME biodiesel

Fig. 11 Direct ammonia utilization pathways for transportation sector.

The following list summarizes some vehicle powering options and potential applications of ammonia:

• Spark ignited ICE,


• Diesel ICE with H2 or diesel “spike”,
• Combustion turbines,
• Gasoline or ethanol mixture ICEs,
• Transformed biogas generators,
• Direct ammonia fuel cells.

For power generation systems, where the storing space is readily accessible, the energy density is not the responsible aspect for
the fuel choice, as the cost per MJ and emission stages are characteristically the critical factors. With the new energy efficient
systems of making ammonia on the cost per MJ basis, ammonia manufactured via renewable energy resources would be com-
petitive with the fossil-based fuels. The toxicity issue is not also as dangerous for power generation methods since the fuel will be
controlled by professionals following well-established handling processes.

2.1.1.3.1 Ammonia in Heating, Ventilation and Air Conditioning applications


Ammonia has been recognized and employed as a leading refrigerant in the industrialized regions due to its outstanding thermal
features, zero ozone depletion and zero global warming potential (GWP). Ammonia has the maximum refrigerating outcome per
unit mass compared to all the refrigerants being used counting the halocarbons. The notable benefits of ammonia over R-134a
could be: inferior overall operational costs of ammonia systems, the flexibility in meeting complex and several refrigeration needs,
and inferior initial costs for several applications [44]. Ammonia is obtainable almost everywhere and is the lowest cost of all the
regularly used refrigerants. Ammonia has superior heat transfer features than most of the chemical refrigerants and consequently
allow for the use of equipment with a smaller heat transfer area. Thus plant building costs will be lower. Furthermore, as these
features also benefit the thermodynamic efficiency in the system, it also diminishes the operational costs of the system. In many
countries, the cost of ammonia per mass is significantly inferior to the cost of HFCs. This kind of advantage is even increased by the
fact that ammonia has a lower density in the liquid phase. Contemporary ammonia systems are entirely closed-loop systems with
completely integrated controls, which adjust the pressures all over the system. Additionally, every refrigeration system is regulated
by codes, which are effective, mature, and continuously updated and revised, to have safety relief valves to protect the system and
its pressure vessels from over-pressurization and possible failure.
For a refrigerant to be considered a long-term option, it is advised to meet three criteria:

• Safe,
• Environmentally friendly,
• Good thermodynamic performance.
Numerous non-halogen materials, containing ammonia, carbon dioxide, and hydrocarbons, work as refrigerants. All of these
materials can be refrigerants for the right use if the system can be planned to meet the main choice criteria. Component and
equipment manufacturers continue to research how these refrigerants perform in systems. Ammonia (NH3) has constantly been a
12 Ammonia

Heating

Condenser

Compressor
Ammonia Electricity
Expansion
power valve
generator

Ammonia
refrigerant

Pump

Evaporator

Liquid separator

Cooling
Fig. 12 Ammonia-based heating, ventilation and air conditioning (HVAC) system schematic.

leading refrigerant in the industrial segment. It is classified as a B2 refrigerant by ASHRAE 34-2013 (Designation and Safety
Classification of Refrigerants) for toxicity and flammability, and therefore governed by strict regulations and codes.
Ammonia is used as refrigerant commonly in the refrigeration structures of food industry like dairies, ice creams plants, frozen
food production plants, cold storage warehouses, processors of fish, poultry and meat, and a number of other uses. Though the
specific volume of ammonia is great, the compressor displacement essential per ton of refrigeration is fairly minor, because small
compressor is desired per ton of the cooling capacity. This saves lots of power in the long run.
For the typical conditions around  151C in the evaporator, the condenser and the evaporator pressures are about 2.37 and
11.67 bar, respectively. Since the pressures are not very high, lightweight substances can be used for the building of the equipment.
The pressure in the evaporator is quite high, so it is not necessary to expand the gas to very low pressure. This also empowers high
suction pressure for the compressor and lower compression ratio. The release temperature of the ammonia refrigerant from the
compressor is high, hence water cooling of the cylinder heads and the cylinders of the compressor is vital. If high discharge
pressure is necessary, it is desirable to use the multi-cylinder compressors instead of the single cylinder compressor to evade
overheating of the compressor.
Recently, alternative ammonia chilled water systems are also developed. One of these examples is the elimination of com-
pressor. Although there are some chilled water systems in residential applications, they are mostly employed in commercial air
conditioning systems.
A basic schematic of heating, ventilation and air conditioning (HVAC) systems is illustrated in Fig. 12. Ammonia can be
used as a refrigerant in the cycle of HVAC systems. Additionally, for stand-alone applications, the power required for a compressor
of HVAC system can be produced by ammonia-based power generation units so called ammonia generators as shown in
Fig. 12.
The compressor sucks the dry gas (from the evaporator and flash gas) from the separator at evaporating temperature, com-
presses it to condensing temperature and feeds the superheated discharge gas to the condenser. The condenser liquefies the
refrigerant while dissipating the heat from the refrigerant gas to the cooling media. From the condenser, the liquid refrigerant is fed
to the expansion device at condensing pressure and close to the condensing temperature. In the expansion device, the ammonia is
expanded to evaporating temperature and then fed to the separator. In the separator, liquid and flash gasses are separated. The
liquid refrigerant, at evaporating temperature and pressure, is sucked by the pump and delivered to the evaporator. In the
evaporator, the heat exchange takes place. A mix of gas and liquid is fed back to the separator, where the liquid is separated from
the gas, and the compressor can suck dry gas.

2.1.1.3.2 Ammonia as both fuel and refrigerant


Ammonia has outstanding potentials as a refrigerant and as a fuel. It is also worth to examine the option to cool the engine with
ammonia that can act as a refrigerant while it is heated to the temperature at which it is fed to the power producer (ICE or fuel
cell). Optionally, the cooling outcome of ammonia, i.e., its high latent heat of evaporation, may be used to harvest some air
Ammonia 13

40

Proposed DOE goal 2020


35

Gasoline
30

Ammonia
Metal amines
25

Chemical hydrides
GJ m−3

Methanol
Liquefied H2
20

Compressed H2

LPG
15

Bateries
10

CNG
5

0
0 10 20 30 40 50 60
GJ t−1
Fig. 13 Comparison of volumetric energy densities and specific energy densities of various fuels and ammonia. Modified from Zamfirescu C,
Dincer I. Ammonia as a green fuel and hydrogen source for vehicular applications. Fuel Process Technol 2009;90:729–37.

Table 4 Conversion properties of hydrogen-fueled ICE Ford Focus to run on


NH3 fuel

Property Unit H2 NH3

Volume of storage tank Liter 217 76


Pressure of storage Bar 345 10
On-board energy MJ 710 1025
Cost of full tank $ 18.87 10.57
Range of drive km 298 430
Cost of drive $ 100 km1 6.34 2.42
Compactness of tank L 100 km1 73 18

Source: Reproduced from Zamfirescu C, Dincer I. Ammonia as a green fuel and hydrogen source for vehicular
applications. Fuel Process Technol 2009;90:729–37.

conditioning onboard. The comparison of volumetric energy densities and specific energy densities of numerous fuels is illustrated
in Fig. 13.
Numerous automakers have industrialized the prototypes of hydrogen-fueled cars in recent years. Here, for examination
purposes, a Ford Focus H2ICE prototype is selected [3]. In Table 4, it is listed the performance parameters of the real prototype and
some calculation results for the similar prototype as converted to NH3 fuel. In calculation it has been assumed that the price of
ammonia is $ 0.23 kg1 and the power-train performance is characterized by 1.19 MJ km1 shaft power where it is founded on
specified 50% efficiency, 710 MJ stored in the full tank and 298 km driving range [3]. The effectiveness of the ammonia engine has
been taken the similar as the hydrogen engine. Actually, ammonia can be dissociated onboard at no extra cost (only using the heat
rejected by the ICE) and the engine fueled with pure hydrogen [3].

2.1.1.3.2.1 Thermo-catalytic decomposition of ammonia


Ammonia can be decomposed thermo-catalytically to generate hydrogen according to the following endothermic reaction [40]:
2 1
NH3 þ 30:1 kJ mol1 H2 -H2 þ N2 ð1Þ
3 3
Here, the required enthalpy signifies 10.6% of HHV or 12.5% of the lower heating value (LHV) of the generated hydrogen. The
ammonia decomposition reaction does not need catalysis to be performed at high temperatures for example over 1000K; though,
at inferior temperatures, the reaction rate is too slow for practical applications such as hydrogen generation for energy conversion.
At 4001C, the equilibrium conversion of NH3 is very high at 99.1% [45] and at about 4301C, almost all ammonia is converted to
hydrogen at equilibrium, below atmospheric pressure circumstances [11]. There is a big array of catalysts appropriate to ammonia
decomposition (e.g., Fe, Ni, Pt, Ir, Pd, and Rh), nonetheless ruthenium (Ru) seems to be the finest one when reinforced with
carbon nanotubes, making hydrogen at additional than 60 kW equal power per kilogram of catalyst [45]. Over ruthenium
catalysts, at temperatures lower than about 3001C, recombination of nitrogen atoms is rate limiting, while at temperatures higher
than 5501C, the cleavage of ammonia’s N–H bond is rate limiting. Though, the activation energy is greater at low temperature
(180 kJ mol1) and inferior at higher temperatures (21 kJ mol1). The finest temperature range for ammonia decomposition
14 Ammonia

over ruthenium catalysts may be 350–5251C, which proposes that flue gases from hydrogen ICEs, other hot exhausts from
burning equipment, or electrochemical power conversion in high-temperature fuel cells can be used to drive ammonia decom-
position [40].

2.1.1.3.3 Ammonia and urea


The mission of finding the optimal hydrogen carrier is not easy as it includes multi-criteria decision making and attention of
numerous practical and financial characteristics of safety, energy density and cost of processing or recycling. This has led to the
inspection of a miscellaneous spectrum of storing resources such as metal hydrides, metal-organic materials, and amide systems
[46]. In spite of wide research and improvement determinations, these equipment and composites have main disadvantages
revolving around the rate of hydrogen desorption, cyclability, and high cost [47,48].
With this respect, ammonia has been regarded as an excellent hydrogen carrier for its several favorable attributes as shown in
Table 5. Large quantities of ammonia are used worldwide for agricultural purposes. The infrastructure and technology of ammonia
production are also well established with existing industrial plants around the world to support the increasing demand for
fertilizers [9]. Natural gas is the main feedstock for the synthesis of ammonia which uses the steam reforming method. So from a
life-cycle perspective, the production of 1 t of ammonia emits about 1.5 metric tons of carbon dioxide most of which can be easily
recovered for use in downstream processes such as the manufacture of urea or other derivatives [49]. This figure excludes the
potential amount of carbon dioxide emitted if carbon-based fuel is used to provide the energy required to drive the process of
ammonia production.
On the other hand, ammonia is corrosive, toxic, and life-threatening when released at high concentrations [50]. To lessen these
risks, some attention has been focused toward steadying the ammonia by merging it in metal ammine complexes or ammonia-
borane systems. This permits for the transportation and long-term storage of fuel in solid state or liquid form and hydrogen can be
released on demand [30]. However, such systems are also burdened with disadvantages like to those discussed earlier. Alter-
natively, urea is a nontoxic chemical which can be found in natural systems as well as human and animal waste (urine). On
average, the concentration of urea in human urine is 9.3–23.3 g L1 [51]. Pure urea is formed as white, odorless prills, or granules
when artificially synthesized. Owing to its stable nature, it can be easily and safely handled, transported and stored at room
temperature. Also, urea is the most widely used solid fertilizer worldwide. In 2009, the global production of urea reached 146
million tons, and it is anticipated to increase to 210 million tons by 2013 due to increasing global demand. This major increase is,
due to the growth of the nonagricultural use of urea in emission control (DeNOx) systems for industrial and automotive
applications [52]. As stated earlier, the process of ammonia production normally supplies the feedstock of ammonia and carbon
dioxide for the synthesis of urea. Therefore, greenhouse gas is released only when fossil fuel is utilized to provide the required
energy for this process.

2.1.2 Analysis and Assessment

In this section, ammonia and urea usage in fuel cell applications is discussed by electrochemical analyses presented before in
Ref. [52]. Ammonia and urea are directly used in SOFCs for power generation.

2.1.2.1 Ammonia in Direct Ammonia Solid Oxide Fuel Cell


When heated, ammonia is decomposed to hydrogen and nitrogen according to the following reaction:

3 1
NH3 2 H2 þ N2 ð2Þ
2 2
The above reaction is endothermic and depends on the temperature and pressure of the system. The reaction proceeds to the
right as the temperature increases or to the left as the pressure increases thus producing less hydrogen. For the purpose of this
analysis, it is assumed that the ammonia is directly fed to the fuel cell where it will be consumed and the only species involved in
the decomposition process are NH3, H2, and N2.
Catalysts are used to lower the activation energy and speed up the rate of ammonia decomposition reaction within the fuel cell.
Much effort has been focused on finding a suitable catalyst to promote the decomposition of ammonia at low temperature and

Table 5 Energy density of different energy carriers (based on LHV value)

Energy carrier Density (kg m3) Gravimetric density (%H2) Volumetric density (kg H2 L1) Energy density (MJ L1)

Gaseous H2 (298K, 10 MPa) 7.68 100 0.0077 0.92


Liquid H2 (30K, 10 MPa) 72.58 100 0.0726 8.71
Liquid NH3 (298K, 1 MPa) 603 17.76 0.1071 12.85
Aqueous urea (76.92%wt – STP) 1200 7.74 0.0930 11.16

Sources: Reproduced from Ma Q, Ma J, Zhou S, et al. A high-performance ammonia-fueled SOFC based on a YSZ thin-film electrolyte. J Power Sources 2007;164 86–9 and Egan EP,
Luff BB. Heat of solution, heat capacity, and density of aqueous urea solutions at 251C. J Chem Eng Data 1966;11:192–4.
Ammonia 15

high kinetic rate. Although ruthenium-based catalysts are expensive, they can achieve near complete decomposition of ammonia at
temperatures around 600K. More attractive alternatives include nickel-based catalysts which are relatively cheaper and can achieve
the same task at about 700–800K [11]. Since the operating temperature of an average SOFC ranges between 900 and 1400K [53], it
is safe to neglect the kinetic rate of reaction and assume that full decomposition of ammonia is attained within the porous
anode layer.

2.1.2.2 Urea in Direct Urea Solid Oxide Fuel Cell


An aqueous solution composed of equimolar amounts of urea and water (76.92% urea by weight) can be heated and injected into
the system resulting in a transitory evaporation of water from the spray droplets such that [52]:
ðNH2 Þ2 COðaqÞ -ðNH2 Þ2 COðsÞ þH2 OðgÞ ð3Þ
The rapid thermolysis of solid urea can be initiated at a relatively low temperature of 406K to yield ammonia and isocyanic acid
as
ðNH2 Þ2 COðsÞ -NH3ðgÞ þHNCOðgÞ ð4Þ
while the simultaneous hydrolysis of isocyanic acid over a catalyst like titanium oxide anatase gives
HNCOðgÞ þ H2 OðgÞ -NH3ðgÞ þ CO2ðgÞ ð5Þ
resulting in the following reaction:
ðNH2 Þ2 COðsÞ þ H2 OðgÞ -2NH3ðgÞ þ CO2ðgÞ ð6Þ
Further heating of the products over a suitable catalyst like nickel will initiate the decomposition of ammonia to hydrogen and
nitrogen to yield the final reaction
ðNH2 Þ2 COðsÞ þ H2 OðgÞ -3H2ðgÞ þ N2ðgÞ þ CO2ðgÞ ð7Þ
Furthermore, other chemical species like carbon (graphite) and carbon monoxide formed during intermediate or side reactions
are considered for determining the thermodynamic equilibrium of the thermohydrolysis of urea.

2.1.2.3 Utilization of Fuel and Oxidant


The utilization of ammonia or urea can be defined in terms of the actual supply and consumption of the fuel or its hydrogen
equivalent such that
Fuelconsumed
Uf ¼ ð8Þ
Fuelsupplied
In the same way, the utilization of air as an oxidant can be witten as a function of the air supply and usage or its oxygen
equivalent because air can be presumed to be composed of 79% nitrogen and 21% oxygen by volume
Air consumed
Uf ¼ ð9Þ
Air supplied
This applies to ion-conducting and proton-conducting cells.

2.1.2.4 Electrochemical Analysis


The examination of SOFCs at closed-circuit conditions requires the characterization of the overpotentials affecting the cell
operation. The working potential or voltage of SOFC can be determined as
V ¼ E  jactan  jactca  jO  jconcan  jconcca ð10Þ
The models described below are applicable to all types of SOFC under investigation in this study.

2.1.2.4.1 Activation overpotential


This kind of irreversible loss is linked to the electrode kinetics and signifies the difference between the real potential (closed-circuit
conditions) and the reversible potential (open-circuit conditions). The implies connection between the current density and the
activation overpotential which can be defined using the Butler–Volmer expression [54,55]:
    
azFjact ð1  aÞzFjact
J ¼ J0 exp  exp  ð11Þ
RT RT
The exchange current density (J0) is a measure of the magnitude of electron activity at the equilibrium potential of the
electrode. The value is greatly inclined by the electrode material, structure and other issues such as the temperature of the reaction
and the length of triple phase boundary (TPB) [56]. The charge transfer coefficient (a) designates the effect of the electrical
potential on the ratio of the forward to the reverse activation blockade. For most electrochemical reactions in fuel cells, this value is
assumed as 0.5 [57,58]. In addition, the parameter (z) refers to the number of electrons transferred per mole of fuel. Therefore, the
16 Ammonia

explicit relationship of the activation overpotential is


!
RT J
jacty ¼ sinh1 ð12Þ
F z J0y

Here, y is anode and cathode.

2.1.2.4.2 Ohmic overpotential


The electronic resistance of the fuel cell electrode and connection is typically subtle and is deliberated negligible when compared to
the ionic resistance of the electrolyte which can be described by Ohm’s law [58,59].
jO ¼ JδRO ð13Þ
The electrolyte resistance (RO) is a function of the electrolyte features and is highly reliant on the operational temperature.

2.1.2.4.3 Concentration overpotential


The concentration overpotential accounts for losses experienced by the resistance of the porous electrode to the carriage of gaseous
species between the gas channel and the reaction sites at TPB. The mass transport in the electrodes is driven by the diffusion of
reacting species because of concentration gradient as well as the infusion triggered by pressure gradient [60,61]. Numerous mass
transport models with changing accuracies have been used to capture the effects of concentration overpotential on the perfor-
mance of fuel cells. It has been demonstrated that the highest accuracy can be achieved using the dusty-gas model (DGM) [60] and
the mean transport pore model (MTPM) [62]. The one-dimensional multi-component mass transport model using DGM can be
written as
( !)
Ni X yi Ni  yi Nj 1 dyi dP B0 P
þ ¼  P þ yi 1þ ð14Þ
Deff
ik
j ¼ 1; ja i Deff
i;j RT dx dx mmix Deff
i;k

where the effective Knudsen diffusion accounts for the porosity and tortuosity of the electrode such that [61]
sffiffiffiffiffiffiffiffiffiffi
2 ϵ 8 RT
Di;k ¼
eff
rp ð15Þ
3 t p Mi
The effective binary diffusion of chemical species can be obtained using the Chapman–Enskog equation [63]
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 
ϵ 1 1 1
Dij ¼ 0:0018583
eff
T3 þ ð16Þ
t Mi Mj P s2ij OD;ij
and the mean characteristic length (sij) of the molecular collision diameter of species i and j is given by
si þ sj
sij ¼ ð17Þ
2
The collision integral (OD,ij) is a function of temperature and the Lennard-Jones parameter (k/eij)
1:06036 0:193 1:03587 1:76474
OD;ij ¼ þ þ þ ð18Þ
G0:15610 e0:47635 G e1:52996 G e3:89411 G
where
kT
G¼ ð19Þ
eij
pffiffiffiffiffiffiffi
eij ¼ ei ej ð20Þ
The values of the parameter in Eqs. (16)–(20) have been obtained from Refs. [63–65]. The permeability of the porous electrode
is estimated using the Kozeny–Carman relationship [66]
4 ϵ3 rp2
B0 ¼ ð21Þ
72t ð1  ϵÞ2
Furthermore, the viscosity of gas mixture can be calculated using the following semi-empirical equation [67]:
P pffiffiffiffiffiffi
yi mi Mi
mmix ¼ P p ffiffiffiffiffiffi ð22Þ
yi Mi
The gas viscosities are obtained from Ref. [63]. As presented by Zhu and Kee [68], the evaluation of the pressure gradient (dP/
dx) across the electrode can be written as
 
P Ni
i ¼ 1 Deff
dP i;k
¼    P   ð23Þ
dx B0 P yi
1
RT
þ mRT i ¼ 1 Deff
i;k
Ammonia 17

At TPB, the molar flux of the gaseous reactants involved in the electrochemical reaction, namely hydrogen and oxygen, can be
related to the current density as [69]
J
Ni ¼ ð24Þ
zF

where z is equal to two for hydrogen and four for oxygen. Depending on the type of electrolyte (SOFC-O or SOFC-H), the molar
flux of water vapor can be calculated using Graham’s law of diffusion which governs the diffusion in gas mixtures [60]:
X pffiffiffiffiffiffi
Ni Mi ¼ 0 ð25Þ
i

The molar fluxes of all other non-reacting species are equal to zero [69]. The set of simultaneous differential equations can be
solved numerically using the software to obtain the molar fractions and partial pressures of the reacting species at TPB. Finally, the
relationship between the concentration overpotential and the partial pressures can be written as [57,70]
For SOFC-O:
!
RT pTPB
H2 p H2 O
jconcan ¼ ln ð26Þ
2F pH2 PH
TPB
2O

!
RT p O2
jconcca ¼ ln TPB ð27Þ
4F P O2

For SOFC-H:
!
RT p H2
jconcan ¼ ln TPB ð28Þ
2F P H2

0 1
TPB pffiffiffiffiffiffiffi
P
RT B H2 O O2 C p
jconcca ¼ ln@ qffiffiffiffiffiffiffiffiffiA ð29Þ
2F PH2 O PO TPB
2

2.1.2.4.4 Electrochemical efficiency


The maximum electrochemical efficiency of SOFC can be defined as [71]
z F E Uf
Zelc ¼ ð30Þ
DH0
Since only hydrogen is oxidized in the fuel cell, its LHV is used in this particular calculation.

2.1.2.5 Illustrative Example


Using the Gibbs energy minimization (GEM) method described earlier, the thermodynamic equilibrium of ammonia is obtained
as depicted in Fig. 14. It can be seen that the decomposition of ammonia is favored at higher temperatures and low pressures.
Nonetheless, it is well known that the reaction kinetics is rather slow and often requires the use of a catalyst to promote faster
conversion of ammonia. This example is derived from Ref. [52].
Complete transformation of ammonia may be realized over expensive metal catalysts such as ruthenium and platinum at
temperatures around 600–650K [72]. However, more economical alternatives include iron, nickel or nickel-based catalysts. Fig. 15
shows the open cell voltage (OCV) of the DA-SOFC as a function of fuel utilization. It is permanently desired to have a high fuel
utilization to maximize the system efficiency and evade fuel combustion and its emissions in combustion chambers and after-
burners. However, higher OCV value is a challenging goal which tends to decay as the fuel utilization is improved. As a general
trend, the OCV of SOFC-O is lower than that of SOFC-H and the difference becomes more pronounced at higher fuel utilizations
due to the high rate of the oxidation reaction and water vapor formation.
The OCV value of DU-SOFC is usually lower than that of DA-SOFC mainly due to the presence of additional chemical species
like carbon dioxide and carbon monoxide which reduce the partial pressure of hydrogen. This is shown in Fig. 16.
Figs. 17 and 18 indicate that the OCV of the corresponding types of SOFC can be improved by 5–6% when the operating
pressure is increased by five times the atmospheric pressure. This modest enhancement is of particular benefit when SOFC is used
in an integrated pressurized system.
Figs. 19 and 20 show the maximum electrochemical efficiency of the respective fuel cells for a given fuel utilization. In general,
the maximum efficiency of SOFC-H is higher than that of SOFC-O due to higher OCV over the range of fuel utilizations. Moreover,
DA-SOFC is more efficient than DU-SOFC by 3–8% at high fuel utilization due to a higher average molar fraction of hydrogen
under most conditions. From the figures, it can also be seen that operating SOFC at higher pressure improves the OCV which in
turn offers a slight enhancement to the maximum electrochemical efficiency. The vertical dashed lines represent the maximum fuel
utilization at which the maximum electrochemical efficiency can be achieved.
18 Ammonia

1.0

0.9
NH3
0.8 H2

0.7

Molar fraction (–)


0.6
101 kPa
0.5 505 kPa
1010 kPa
0.4

0.3 N2

0.2

0.1

0.0
300 400 500 600 700 800 900 1000 1100 1200 1300
Temperature (k)
Fig. 14 Thermodynamic equilibrium of ammonia. Data from Ishak F. Thermodynamic analysis of ammonia and urea fed solid oxide fuel cells
(SOFCs); 2011.

1.1
SOFC-H

1.0

0.9
OCV (V)

SOFC-O
0.8
P = 101 kPa
1073K
Air utilization = 25% 1273K
0.7

0.6
0 20 40 60 80 100
Ammonia utilization (%)
Fig. 15 Open cell voltage (OCV) of DA-SOFC at different temperatures. Data from Ishak F. Thermodynamic analysis of ammonia and urea fed
solid oxide fuel cells; 2011.

The following concluding remarks can be derived from the illustrative example:

• The proton-conducting ammonia fed SOFC maintained the highest reversible cell potential under all open-circuit conditions
followed by the ion conducting counterpart. The ion and proton-conducting urea fed fuel cells attained lower cell potentials
due to higher overpotentials
• Under closed-circuit conditions, the proton-conducting ammonia fed SOFC demonstrated the best performance due to the
higher partial pressure of hydrogen at the anode in comparison to the ion-conducting counterpart. In addition, intermediate
and side reactions at the anode of ion and proton conducting urea fed SOFC resulted in a strong decrease of the hydrogen
partial pressure hence the lower performance.
• The reverse water gas shift reaction has a harmful consequence on the partial pressure of hydrogen at the anode of urea fed
SOFC. Especially, the effect is most marked in the proton-conducting SOFC primarily due to the nonappearance of sufficient
water vapor at the anode of the cell.
Ammonia 19

1.1
SOFC-H

1.0

0.9

OCV (V)
0.8 SOFC-O

P = 101 kPa
1073K
0.7 SUR=1
1273K
Air utilization = 25%

0.6
0 20 40 60 80 100
Urea utilization (%)
Fig. 16 Open cell voltage (OCV) of DU-SOFC at different temperatures. Data from Ishak F. Thermodynamic analysis of ammonia and urea fed
solid oxide fuel cells (SOFCs); 2011.

1.1
SOFC-H

1.0

0.9
OCV (V)

SOFC-O
0.8

T = 1073K P = 101 kPa


0.7 Air utilization = 25% P = 505 kPa

0.6
0 20 40 60 80 100
Ammonia utilization (%)
Fig. 17 Effect of pressure on the open cell voltage (OCV) of DA-SOFC. Data from Ishak F. Thermodynamic analysis of ammonia and urea fed
solid oxide fuel cells (SOFCs); 2011.

1.1
SOFC-H

1.0

0.9
OCV (V)

SOFC-O

0.8
T = 1073K 101 kPa
0.7 SUR=1 505 kPa
Air utilization = 25%

0.6
0 20 40 60 80 100
Urea utilization (%)
Fig. 18 Effect of pressure on the open cell voltage (OCV) of DU-SOFC. Data from Ishak F. Thermodynamic analysis of ammonia and urea fed
solid oxide fuel cells (SOFCs); 2011.
20 Ammonia

100

Maximum electrochemical efficiency (%)


P = 101 kPa
90 P = 505 kPa
80 SOFC-H
70
60
50
40 SOFC-O
30
20
T = 1073K
10
Air utilization = 25%
0
0 10 20 30 40 50 60 70 80 90 100
Ammonia utilization (%)
Fig. 19 Maximum electrochemical efficiency of DA-SOFC. Data from Ishak F. Thermodynamic analysis of ammonia and urea fed solid oxide fuel
cells (SOFCs); 2011.

100
Maximum electrochemical efficiency (%)

90 P = 101 kPa
P = 505 kPa
80
70 SOFC-H

60
50
40 SOFC-O
30
20
T = 1073K
10 SUR = 1
Air utilization = 25%
0
0 10 20 30 40 50 60 70 80 90 100
Urea utilization (%)
Fig. 20 Maximum electrochemical efficiency of DU-SOFC. Data from Ishak F. Thermodynamic analysis of ammonia and urea fed solid oxide fuel
cells (SOFCs); 2011.

2.1.3 Case Study 1: Ammonia Utilization in Road Vehicles

In this case study, a comparative life cycle assessment (LCA) of ICE-based vehicles fueled by various fuels, ranging from hydrogen to
gasoline, is conducted in addition to electric and hybrid electric vehicles. Three types of vehicles are considered, such as ICE vehicles
using gasoline, diesel, LPG, hydrogen, and ammonia; hybrid electric vehicles using gasoline and electricity; and electric only vehicles
for comprehensive comparison and environmental impact assessment. The processes are analyzed from raw material extraction to
vehicle disposal using LCA methodology. In order to reflect the sustainability of the vehicles, seven different environmental impact
categories are considered: abiotic depletion, acidification, eutrophication, global warming, human toxicity, ozone layer depletion,
and terrestrial ecotoxicity. The energy resources are chosen mainly conventional and currently utilized options to indicate the actual
performances of the vehicles. The results show that electric and hybrid electric vehicles result in higher human toxicity, terrestrial
ecotoxicity, and acidification values because of manufacturing and maintenance phases. In contrast, hydrogen vehicles yield the most
environmentally benign option because of high energy density and low energy consumption during operation.

2.1.3.1 Materials and Methods


A typical life cycle of a vehicle technology can be categorized into two main steps, namely fuel cycle and vehicle cycle. In the fuel
cycle, the processes beginning from the feedstock production to fuel utilization in the vehicle are considered. The extraction of
crude petroleum is accounted for diesel. Transformation of crude oil feedstock into useful fuels is a too energy intensive stage of
Ammonia 21

Vehicle maintenance

Vehicle
Vehicle disposal Vehicle operation manufacturing
Transport services
Transport services Energy production
Energy production
Infrastructure Raw material

Production/assembly/
Energy production Energy Energy consumption infrastructure and auxiliaries
transmission
Fig. 21 Boundaries of conducted LCA analyses including fuel and vehicle cycle.

the fuel cycle, producing substantial amounts of GHG. Nevertheless, purification of natural gas results in considerably fewer energy
usage and GHG. The fuel needs to be transported to be available for vehicle usage. Hence, emissions and energy usages associated
with fueling trucks/pipelines are thus accounted for in the fuel delivery step [73]. The boundaries of the current analyses are
illustrated in Fig. 21 including fuel and vehicle cycles.
A remarkable part of any life cycle analysis involves gathering of reliable data. The excellence of data has a deep influence on the
quality of the results predicted or estimated by an LCA tool. Argonne National Laboratory has developed a full life cycle model called
GREET (Greenhouse gasses, Regulated Emissions, and Energy use in Transportation), sponsored by the U.S. Department of Energy,
Office of Energy Efficiency and Renewable Energy, which allows evaluating various vehicle and fuel combinations on a full fuel cycle
or vehicle cycle basis. It is a powerful software with a substantial amount of alternative vehicles. It takes into account the well to
pump (WTP) and well to wheel (WTW) processes. WTW phases for each fuel starts with the extraction of the primary energy and ends
with the consumption in vehicles. WTP consists of the feedstock and fuel phases, counting fuel feedstock extraction, transmission,
distribution, and storage. Pump to the wheel (PTW) represents the energy use and emissions during vehicle operation. The functional
unit is the distance traveled by the vehicle which is taken as 1 km for each type of vehicle. The results of the primary energy demand
and the GHG emissions for each vehicle type are expressed as MJ km1 and g CO2 eq. km1, respectively [67].
In this case study, the GREET 2015 software is utilized to simulate the life cycle GHG emissions for the selected vehicles only
for the operation process. After the required operation data is obtained from the GREET software, they are used in the SimaPro
LCA tool for complete LCA analyses. The scope of the analyses represents a complete LCA since they include the WTW stages as
well as the equipment life cycle. The equipment life cycle includes production, manufacturing, maintenance, and end-of-life of
vehicle infrastructure. the GREET software can calculate energy use with high accuracy. In this case study, the vehicle operation
emissions belonging to different types of vehicles and transportation fuels are based on GREET 2015 calculations [74]. A part of
GREET model characterizes the life cycle of vehicles, including production, maintenance, operation, and disposal [75]. In parti-
cular, electric and hybrid electric vehicle data are utilized from production until disposal and then adapted in the SimaPro LCA
software. The LCA database ecoinvent, v2.2 was used as source of background LCI data. Life cycle impact assessment (LCIA),
quantification of life-cycle environmental weights and potential impacts was carried out using the LCA software SimaPro [76].
Impact assessment stage in LCA study is the part in which collected inputs and outputs of fundamental streams are interpreted
into impact pointer results typically linked to human health, environment, and resource depletion. It is essential to remind that
LCA and impact assessment analyses indicate the possible environmental impacts formed by exchanges that cross the border
between technosphere and ecosphere, and act on the natural environment and humans. The results of LCIA are understood as
environmentally suitable indicators of potential impact, rather than estimates of real environmental effects.
A few standards from the International Organization for Standards (ISO) such as ISO 14040 and ISO 14044 govern the exact
requirements necessary to manage LCA studies [77,78]. A guide for the operation of ISO standards was published by the Center of
Environmental Science of Leiden University (CML) in 2001 [79]. This guide defines the process to be realized for studying LCA
project agreeing to the ISO standards. In the current study, a CML impact assessment methodology is applied. In the impact
assessment phase of LCA, a group of impact categories and the characterization methods and factors for a wide list of materials are
suggested. For applying the structures in the ecoinvent life cycle inventory database, it is essential to allocate the characterization
factors to the elementary source streams and pollutant streams described in the database. The categories considered in this case
study are shortly described here.
Human toxicity: Toxic substances on the human environment are the main concerns for this category. 1,4-Dichlorobenzene
equivalents kg1 emissions are used to express each toxic substance. Reliant on the material, the geographical scale varies between
local and global indicator.
Global warming: The greenhouse gasses to air are related to the climate change. Adverse effects upon ecosystem health, human
health, and material welfare can result from climate change. A kg carbon dioxide per kg emission is used to express the GWP for
time horizon 500 years (GWP500). It has a global scale.
Acidification potential: Acidifying substances causes a broad range of impacts on soil, groundwater, surface water, organisms,
ecosystems, and materials. SO2 equivalents kg1 emission is used to expresses the acidification potential.
22 Ammonia

Eutrophication: This category considers the impacts of to extreme intensities of macro-nutrients in the environment initiated by
emissions of nutrients to air, water, and soil. It is stated as kg PO4 equivalents per kg emission and the terrestrial measure differs
between local and continental scale, the time span is infinity.
Depletion of abiotic resources: This impact group is related to protection of human well-being, human health, and ecology
health. The indicator is connected to the extraction of minerals and fossil fuels because of involvements in the structure. It is
recognized for every extraction of minerals and fossil fuels depending on the concentration of resources and rate of deaccumu-
lation. The terrestrial scope of this indicator is provided at a global scale. It is expressed in kg antimony equivalents/kg extraction
unit.
Stratospheric ozone depletion: A superior portion of UV-B radiation spreads the earth surface because of increasing chloro-
fluorocarbons (CFCs). This may yield damaging impacts on human and animal health, terrestrial and aquatic ecosystems, and
biochemical cycles and on substances. Ozone depletion potential of several gasses is specified in kg CFC-11 equivalent per kg
emission where the time span is infinity.
Terrestrial ecotoxicity: This category denotes to influences of toxic substances on terrestrial ecosystems. Ecotoxicity potential is
considered with describing fate, exposure and effects of toxic materials. The time horizon is infinite. The results are stated as 1,4-
dichlorobenzene equivalents per kg emission. The indicator relates at global/continental/regional and local scale.
Marine aquatic eco-toxicity: Marine eco-toxicity refers to impacts of toxic substances on marine aquatic ecosystems. It considers
each substance emitted to the air, water or/and soil. The unit of this factor is kg of 1,4-dichlorobenzene equivalents (1,4-DB eq.)
per kg of emission.
Marine sediment ecotoxicity: Marine sediment ecotoxicity refers to impacts of toxic substances on marine sediment ecosystems.
The unit of this indicator is kg of 1,4-dichlorobenzene equivalents (1,4-DB eq.) per kg of emission.

2.1.3.2 Systems Description


On average, lifetime performance of a passenger car is assumed to be 239,000 person km. The average utilization factor is expected
to be 1.59–1.6 passengers per car. Henceforth, the lifetime of the selected vehicles is approximately 150,000 km. Here, the
assessment comprises the following life cycle phases in which the functional unit is 1 km distance traveled.

• Manufacturing of the vehicle,


• Operation of the vehicle,
• Maintenance of the vehicle,
• Disposal of the vehicle.

2.1.3.2.1 Passenger car manufacturing


The inventory contains processes of energy, water, and material usage in passenger car manufacturing. Rail and road transport of
materials are accounted for. The groundwork of plant is involved together with the issues such as land use, building, road, and
parking structure. The material consumption reflects a modern vehicle. The data for vehicle production are representative for
manufacturing sites with an environmental management system. Thus, the resulting data may be an underestimation of envir-
onmental impacts of an average vehicle fleet.
The electricity comes from a mixture of Union for the Co‐ordination of Transmission of Electricity (UCTE) countries. The UCTE
is the association of transmission system operators in continental Europe in which about 450 million people are supplied
electricity. In the electricity usage process, electricity production in UCTE, the transmission network and direct SF6-emissions to air
are included. Electricity losses during medium-voltage transmission and transformation from high-voltage are also accounted for.
The conversion of high–medium voltage as well as the transmission of electricity at medium voltage are taken into account. In
particular, the required high temperature is assumed to be from methane burned in industrial furnace greater than 100 kW which
in turn contains fuel feed from the high-pressure gas network, infrastructure (boiler), emissions, and required electricity of
operation.
In an electric vehicle, the motion is attained from an electric motor and whole energy used for traction is kept in a battery
system. The car can trip if sufficient energy is available in the battery. When the battery energy is consumed, the battery needs
recharging by electricity grid or replacement. The operation of EV varies from the conventional vehicles in some characteristics as
the first difference is the energy source for operation where electricity is utilized despite petrol or diesel. Henceforth, there are no
tailpipe emissions. It is therefore assumed that emissions are restricted to tire and brake wear and abrasion from the surface of the
road. Conceptually, a HEV is very parallel to EV with the exception that it involves a fuel tank and ICE. Whenever the battery
energy is consumed, the ICE can be used to recharge the battery or for traction power. In the current study, HEV is assumed to be
50% electric and 50% gasoline. For electric and hybrid electric vehicles, the required amount of steel is lower compared to
conventional cars since there is no ICE in the car. However, for EVs and HEVs, production of electric motor and lithium-ion
batteries are included. The battery is the on-board energy storage system of the car. It contains an array of connected cells, the
packaging, and the battery management system. If a car is intended to have a range of 100 km with a regular consumption of about
0.2 kWh km1, a minimum of 20 kWh of energy needs to be stored on-board. When energy density of 0.1 kWh kg1 is assumed
for the battery, the total battery weight would be about 250 kg. Currently, commercially available electric car batteries range
between 100 and 400 kg, depending on automobile size and preferred range [80]. The average masses of an electric motor for and
Ammonia 23

lithium-ion battery are assumed to be 104 kg and 312 kg, respectively, for this case study [81]. The inventory data for battery
production and disposal is utilized from GREET model [74] and [82].

2.1.3.2.2 Maintenance
The inventory of maintenance of vehicles contains resources used for alternation parts and energy consumption of garages. Rail
and road transportation of supplies are accounted for. For EVs, during the lifetime of the car, one battery change is assumed.
Henceforth, lithium-ion battery replacement and disposal processes are also taken into account in the maintenance phase.

2.1.3.2.3 Disposal of the vehicles


The inventory of vehicle disposal contains disposal processes for bulk materials. For the disposal of tires, a cut off allocation is
applied. In addition, the transportation of tires to the cement works is taken into account. For the disposal of steel, aluminum,
copper and tires, a cut off allocation is applied. Waste specific water together with air emissions from incineration and supple-
mentary supply depletion for flue gas scrubbing are accounted for. Short term releases to river water and long-term emissions to
groundwater from slag section and remaining material landfill are considered with process energy loads for municipal waste
incineration plant (MSWI). The following processes are applied for the disposal of a vehicle scenario:

• Disposal of plastics in mixture with 15.3% water to municipal incineration (65 kg). Energy production net output. The waste
yields 0.01693 kg of slag and 0.006594 kg of remains per kg of waste. They are landfilled. Supplementary solidifying with
0.002638 kg of cement is applied. MSWI: 3.48MJ kg1 waste energy in electric form and 7.03MJ kg1 waste energy in thermal
form.
• Disposal of glass to municipal incineration (30.1 kg). Energy production in MSWI net output: 3.67MJ kg1 waste electric energy
and 7.39MJ kg1 waste thermal energy. The waste yields 0.01704 kg of slag and 0.01217 kg of residues per kg of waste. They are
landfilled. Supplementary solidifying with 0.004869 kg of cement is applied.
• Disposal of emulsion paint leftovers to HWI (100 kg). Energy production in hazardous waste incineration (HWI) plant, net
output: 17.11 MJ kg1 electric energy and 1.27MJ kg1 thermal energy. The waste yields 0.707 kg of remains per kg of waste.
They are landfilled. Supplementary solidifying with 0.2828 kg of cement is applied.
• Disposal of zinc in car shredder remains to MSWI (5.89 kg). One kg of this waste produces 0.6244 kg of slag and 0.6202 kg of
residues. They are landfilled. Supplementary solidifying with 0.2481 kg of cement is applied.
Note that lithium-ion batteries are recycled for many purposes. The most noticeable one is the retrieval of valued materials and
to follow to ecological laws. Numerous methods are present for recycling lithium-ion batteries with diverse environmental
consequences. Usually, battery recycling procedures can be expressed in three main categories: mechanical, pyrometallurgical, and
hydrometallurgical processes. Hydrometallurgical processes are evaluated to require considerably lesser energy desires compared
to pyrometallurgical processes. In this case study, the hydrometallurgical process for disposal of batteries is selected with an
average efficiency of 57.5% and energy use of 140 kWh t1 [80]. The inventory data for the disposal of batteries are taken from
[80]. For ammonia and hydrogen fueled vehicles, required amount of steel and electrical energy is a little higher than other cars
because of storage tank infrastructure.

2.1.3.2.4 Operation of vehicles


The operation process of the vehicles is one of the key sections of life cycle analyses. In this phase, fuel consumption is involved.
Direct airborne emissions of gaseous materials, particulate matters, and heavy metals are accounted for. Particulate emissions cover
exhaust- and abrasions emissions. Hydrocarbon emissions contain evaporation. Heavy metal emissions to soil and water pro-
duced by tire abrasion are accounted for. The values are based on the operation of an average passenger car. The specific conditions
for the selected vehicles are presented herein:

• Gasoline: All processes on the refinery site excluding the emissions from combustion facilities, including waste water treatment,
process emissions and direct discharges to rivers are accounted for. The inventory data also includes the distribution of
petroleum product to the final consumer including all necessary transports. Transportation of product from the refinery to the
end user is considered together with the operation of storage tanks and petrol stations. Emissions from evaporation and
treatment of effluents are accounted for. Particulate emissions cover exhaust- and abrasions emissions.
• Diesel: Diesel is evaluated as low-sulfur at local storage with an estimation for the total conversion of refinery production to
low-Sulfur diesel. An additional energy use (6% of energy consumption for diesel production in the refinery) has been
estimated. The other processes are similar to gasoline. Particulate emissions cover exhaust- and abrasions emissions.
• Hydrogen: Hydrogen is produced during cracking of hydrocarbons. It includes combined data for all processes from raw
material extraction until delivery at the plant. The output fractions from an oil refinery are composite combinations of mainly
unreactive saturated hydrocarbons. The first processing step in converting such elements into feedstock suitable for the pet-
rochemical industries is cracking. Essentially a cracker achieves two tasks in (i) raising the complexity of the feed mixture into a
smaller number of low molecular mass hydrocarbons and (ii) presenting unsaturation into the hydrocarbons to enable more
reactivity. The raw hydrocarbon input from the refinery is fed to the heater unit where the temperature is increased. The forming
reaction products vary based on the composition of the contribution, the temperature of the heater and the residence time. The
cracker operator selects temperature and residence time to enhance product mix from a supplied input. Cracker feeds can be
24 Ammonia

Table 6 Energy consumptions per km for the selected vehicles

Fuel Fuel/energy consumption Unit

Gasoline 0.0649108 kg km1


Diesel 0.0551536 kg km1
Hydrogen 0.0195508 kg km1
Ammonia 0.0926600 kg km1
Electric vehicle 0.2167432 kWh km1
Hybrid electric vehicle
Electric 0.1083716 kWh km1
Gasoline 0.0324554 kg km1
Liquefied petroleum gas 0.0576296 kg km1

Source: Data from GREET 2015. Argonne, IL: Argonne National Laboratory; 2015.

naphtha from oil refining or natural gas or a mixture of both. After exiting the heater, the hydrocarbon gas is cooled to prevent
extra reactions. After that, it is sent to the separation phase where the individual hydrocarbons are separated from one another
by fractional distillation. Particulate emissions cover exhaust- and abrasions emissions.
• Ammonia: The ammonia synthesis process is Haber–Bosch which is the most common method in the world. Ammonia
production requires nitrogen and hydrogen. In this case study, hydrogen is assumed to be from hydrocarbon cracking as
explained in the previous paragraph. Cryogenic air separation is mostly used method for the massive amount of nitrogen
production. In the LCA of nitrogen production, electricity for the process, cooling water, waste heat and infrastructure for the air
separation plant are included. The Haber–Bosch process is employed in this case study. The reaction is facilitated by catalyst (iron-
oxide based), and the optimal temperature range is 450–6001C. Particulate emissions cover exhaust- and abrasions emissions.
• EV: Electricity consumption is included. Particulate emissions comprise exhaust and abrasions emissions. Heavy metal emis-
sions to soil and water caused by tire abrasion are accounted for. In the electricity usage process, electricity production mix, the
transmission network and direct SF6-emissions to air are included.
• HEV: Hybrid car is assumed to be 50% electric and 50% gasoline with ICE. Electricity and gasoline consumptions are included.
Particulate emissions comprise exhaust and abrasions emissions. Heavy metal emissions to soil and water caused by tire
abrasion are accounted for.
• LPG: All processes on the refinery site excluding the emissions from combustion facilities, including waste water treatment,
process emissions and direct discharges to rivers are considered. All flows of materials and energy due to the throughput of 1 kg
crude oil in the refinery is accounted for. Refinery data include desalting, distillation (vacuum and atmospheric), and
hydrotreating operations. Particulate emissions cover exhaust- and abrasions emissions.
The following fuel consumption rates are considered in the analyses as tabulated in Table 6.

2.1.3.3 Case Study 1 Results and Discussion


The selected vehicle types are environmentally assessed in the SimaPro LCA software based on the energy consumption, and GHG
emissions of ICEV and EVs obtained from the wheel to wheel simulations using the GREET 2015 model. The results presented here
are given on per traveled km basis. The analyzed impact categories are human toxicity, ozone layer depletion, terrestrial ecotoxicity,
eutrophication, acidification, and global warming. The energy use is based on units of MJ. The GWP is presented in kg per CO2
equivalent.
Fig. 22(A) illustrates the total human toxicity values of all vehicles regarding kg 1,4-DB eq. per travel km. EVs and HEVs have
highest human toxicity values corresponding to 0.26 and 0.14 kg 1,4-DB eq. km1, respectively. Compared to other vehicles, they
yield quite higher values because of mainly production and disposal of batteries as shown in Fig. 22(B).
The battery production and assembly requires high amounts of copper and aluminum. Henceforth, top two processes con-
tributing issues related human toxicity are copper and aluminum production at plants. The operation of EVs causes only 1% of
total human toxicity values as shown in Fig. 22(B). HEVs are quite similar to EVs regarding battery manufacturing which yields
second highest value.
The depletion of ozone layer is one of the main reasons for climate change which is actually caused by carbon emissions to the
atmosphere. Since diesel, gasoline and LPG fuels are fossil based and have a huge amount of carbon substance, they have higher
ozone layer depletion values.
The highest is equal to 3.3  108 kg CFC-11 eq. km1 for diesel vehicle as Fig. 22(C) represents. The lowest contributions are
from ammonia and hydrogen vehicle corresponding to 7.19  109 and 1.48  109 kg CFC-11 eq. km1, respectively. Because,
there is no direct CO2 emission during operation of ammonia and hydrogen vehicles.
In parallel to human toxicity values, EVs and HEVs have a higher environmental impact regarding terrestrial ecotoxicity with
0.00026 and 0.00016 kg 1,4-DB eq. km1, respectively, because of copper and steel used in the production process as presented in
Fig. 23(A). For ammonia vehicle, the highest contributor is steel production process together with lignite and heavy oil burned in
power plant.
Ammonia 25

0.3

(kg 1,4-DB eq. km–1)


Human toxicity 500a
0.25
0.2
0.15
0.1
0.05
0
Hydrogen Diesel LPG Gasoline Ammonia Hybrid Electric
(A) vehicle vehicle vehicle vehicle vehicle electric vehicle
vehicle

1% 1% 5% Copper, primary, at refinery


2%
3% Aluminium, primary, liquid, at plant
4%
Anode, aluminium electrolysis
5% Ferrochromium, high-carbon, 68% Cr,
at plant
6%
Disposal, uranium tailings, non-radioactive
emissions
73% Copper, primary, at refinery
Disposal, sulfidic tailings, off-site

(B) Electric vehicle-operation


Ozone layer depletion

(kg CFC-11 eq. km–1)

3.50E−08
3.00E−08
steady state

2.50E−08
2.00E−08
1.50E−08
1.00E−08
5.00E−09
0.00E+00
Hydrogen

Ammonia

Electric

electric

vehicle

Gasoline

vehicle
vehicle
vehicle

Diesel
Hybrid

vehicle
LPG
vehicle

vehicle

(C)

Fig. 22 (A) Life cycle comparison of human toxicity results for various vehicles, (B) contribution of different processes to human toxicity values
of electric vehicles, and (C) life cycle comparison of ozone layer depletion results for various vehicles.

The eutrophication is the impact of excessive levels of macro-nutrients in the environment which is mainly caused by disposal
processes. For EVs, HEVs and ammonia vehicles, the main reason of eutrophication is the disposal of spoil from lignite mining in a
surface landfill in which it corresponds to about 66%, 49% and 47% for ammonia, EVs and HEVs, respectively. The lowest
eutrophication value is observed in hydrogen vehicle with an amount of 7.29  105 kg PO4 eq. km1 as shown in Fig. 23(B).
The acidification values of EVs and HEVs are mainly caused by SO2 emission which corresponds to 70% of overall acidification
value. The source of SO2 emission is predominantly the lignite and bituminous coal at mine that is used for electricity production
mix and eventually consumed in the EVs and HEVs. Afterward, ammonia vehicle has an acidification value of 0.001 kg SO2 eq.
km1 as Fig. 24(A) specifies. 37% of this value is originated from ammonia vehicle operation, and 9% comes from hydrogen
production for ammonia synthesis as shown in Fig. 24(B).
The GWPs of assessed vehicles are comparatively shown in Fig. 25. The lowest GHG emissions are observed in hydrogen,
electric and ammonia vehicles corresponding to 0.049, 0.15 and 0.17 kg CO2 eq. km1, respectively. Hydrogen consumption is
quite lower than ammonia consumption in the passenger car because of higher energy density. It is an expectable result that EVs
also yield lower GWP. However production pathway of electricity has a vital role in GHG emissions. If renewable sources can
realize electricity production, total emissions would decrease for both EVs and HEVs. In addition, the hydrocarbon route for
ammonia vehicle rises the environmental emissions. However, it is noted that renewable-based ammonia is even better than EVs
for most of the impact categories.
26 Ammonia

Terrestrial ecotoxicity 500a (kg 1,4-DB eq. km–1)


0.0003

0.00025

0.0002

0.00015

0.0001

0.00005

0
Electric Ammonia Hybrid Gasoline LPG Diesel Hydrogen
vehicle vehicle electric vehicle vehicle vehicle vehicle
(A) vehicle

Electric vehicle

Ammonia vehicle

Hybrid electric vehicle

Gasoline vehicle

LPG vehicle

Diesel vehicle

Hydrogen vehicle
4

04

04

04

04

4
00

0
E−

E−

E−

E−

E−

E−

E−

E−
E+

00

00

00

00

00

00

00

00
00

1.

2.

3.

4.

5.

6.

7.

8.
0.

(B) Eutrophication (kg PO4 eq. km)–1

Gasoline vehicle

Ammonia vehicle

Hybrid electric vehicle

LPG vehicle

Diesel vehicle

Electric vehicle

Hydrogen vehicle
08

12

14

16

18

2
0

06
0

00

00
00

00

00

00

00

00

00

00
0.

0.
0.

0.

0.

0.

0.

0.

0.

0.

(C) Abiotic depletion (kg Sb eq. km–1)


Fig. 23 (A) Life cycle comparison of terrestrial ecotoxicity results for various vehicles, (B) life cycle comparison of eutrophication results for
various vehicles, and (C) life cycle comparison of abiotic depletion results for various vehicles.
Ammonia 27

Hybrid electric vehicle

Electric vehicle

Ammonia vehicle

Gasoline vehicle

LPG vehicle

Diesel vehicle

Hydrogen vehicle

02

04

06

08

12

14

16
00
00

00

00

00

00

00

00
0.
0.

0.

0.

0.

0.

0.

0.
(A) Acidification (kg SO2 eq. km–1)

4% Ammonia vehicle−operation-
7% hydrocarbon cracking

9% Hard coal, burned in power plant


37%
Lignite, burned in power plant

Hydrogen, cracking, APME, at


20% plant

Heavy fuel oil, burned in power


plant
23%
Operation, transoceanic freight
(B) ship

Fig. 24 (A) Life cycle comparison of acidification results for various vehicles and (B) contribution of different processes to acidification values of
ammonia vehicles.

Hydrogen vehicle

Electric vehicle

Ammonia vehicle

Diesel vehicle

LPG vehicle

Hybrid electric vehicle

Gasoline vehicle

0 0.05 0.1 0.15 0.2 0.25 0.3

Global warming 500a (kg CO2 eq. km–1)

Fig. 25 Life cycle comparison of global warming results for various vehicles.

As Fig. 26 shows, the abiotic depletion is mainly caused by the operation processes of vehicles. Manufacturing, maintenance,
and disposal of EVs have higher shares during life cycle primarily initiated by production and disposal of lithium-ion
batteries.
28 Ammonia

0.0018
Operation Manufacturing, maintenance, disposal
0.0016

Abiotic depletion (kg Sb eq. km–1)


0.0014

0.0012

0.001

0.0008

0.0006

0.0004

0.0002

0
Ammonia Diesel Electric Gasoline Hybrid Hydrogen LPG
vehicle vehicle vehicle vehicle EV vehicle vehicle
Fig. 26 Contribution of operation, manufacturing, maintenance, and disposal processes of the vehicles to abiotic depletion.

Operation
0.25
Global warming 500a (kg CO2 eq. km–1)

Manufacturing, maintenance,
disposal
0.2

0.15

0.1

0.05

0
Ammonia Diesel Electric Gasoline Hybrid Hydrogen LPG
vehicle vehicle vehicle vehicle electric vehicle vehicle
vehicle
Fig. 27 Contribution of operation, manufacturing, maintenance, and disposal processes to overall global warming potential.

On-board storage of hydrogen requires high resistant and strength tanks which lead to higher steel and process requirement.
Henceforth, non-operation part of hydrogen vehicle constitutes about 22 and 44% of overall hydrogen vehicle life cycle for abiotic
depletion and GWP, respectively as illustrated in Figs. 26 and 27. Overall, the operation of the vehicles is dominant contributors to
complete life cycle.

2.1.3.4 Case Study 1 Conclusions


Researchers in recent years intensively investigate alternative fuels for transportation sector. The main criteria for a sustainable fuel
are being environmentally friendly and profitable. In this case study, a comparative environmental impact assessment of alter-
native and conventional fueled vehicles is conducted using the cradle to grave approach via life cycle analyses under different
environmental impact categories. Conventional vehicles considered in this case study include diesel, gasoline, LPG. Alternative
vehicles comprise hydrogen, ammonia, EV and HEV. The analyses are conducted from manufacturing of passenger cars to disposal
including the operation of the vehicles. The results show that hydrogen vehicle is the most environmentally benign one in all
environmental impact categories. Ammonia as a sustainable and clean fuel has lowest GWP after EVs and yield lower ozone layer
depletion values than EVs. Although EVs do not emit direct CO2 during operation, the production and disposal processes of
batteries bring some consequences which harm the environment regarding acidification, eutrophication, and human toxicity. It is
concluded that to have sustainable and clean transportation, production pathway of vehicles, batteries, and alternative fuels need
to be environmentally friendly.
Ammonia 29

2.1.4 Case Study 2: Ammonia Utilization in Maritime Applications

Sea transportation constitutes a large share of global transportation. It is principally used for the transportation of goods, liquid
fuels, all types of products and humans. Transoceanic freight ships need a great amount of energy for an operation which is
commonly provided by diesel or heavy fuel oils. In order to reduce the total greenhouse gas emissions caused by maritime
transportation, alternative fuels, such as ammonia are a potential replacement and/or supplements for conventional fuels. In this
case study, zero carbon fuel – ammonia – is proposed to replace heavy fuel oils in the engines of maritime transportation vehicles.
Furthermore, it is also proposed to use ammonia as dual fuels to quantify the total reduction of greenhouse gas emissions. An
environmental impact assessment of transoceanic freight ship is implemented to explore the impacts of fuel substituting on the
environment. In the life cycle analyses, the complete transport life cycle is taken into account from the manufacture of transoceanic
freight ship to production, transportation and utilization of ammonia in the maritime vehicles. Several ammonia production
routes ranging from municipal waste to geothermal options are considered to evaluate environmentally benign methods com-
paratively. Besides GWP, environmental impact categories of marine sediment ecotoxicity and marine aquatic ecotoxicity are also
selected to examine the diverse effects on marine environment. Being carbon-neutral fuel, ammonia yields significantly minor
global warming impacts during operation. The ecotoxicity impacts on maritime environment vary based on the production route
of ammonia. The results imply that even ammonia is utilized as dual fuel in the engines, the GWP is quite lower in comparison
with heavy fuel oil driven transoceanic ships.

2.1.4.1 Introduction
Decreasing the GWP caused by current transportation technologies and fuels can be reduced significantly by replacing alternative
clean fuels. Sea transportation vehicles mostly use heavy fuel oil or diesel fuel for power generation. Ocean freight ships require a
massive amount of energy for operation. Ammonia is considered as alternative fuel for power generation especially in trans-
portation sector such as maritime. The usage of ammonia in the maritime applications eventually depends on the capability of
producing clean, low-cost energy. One of the hydrogen carriers is ammonia which is synthesized from nitrogen and hydrogen that
can be used for direct combustion in maritime vehicles. Besides having significant advantages in storing and transporting com-
pared to hydrogen, ammonia may also be burned directly in diesel engines. Ammonia can be easily transferred through pipelines,
railway, and ships by delivering to consumption area where it may be utilized as a source of hydrogen, chemical substance, and
fertilizer for agriculture, fuel for transportation such as maritime applications. Since ammonia produces mainly water and nitrogen
on combustion, replacing a part of conventional fuel with ammonia will have a large effect in reducing carbon dioxide emissions.
Ammonia (NH3) is colorless, pungent gas composed of nitrogen and hydrogen. It is the simplest stable compound of these
elements and serves as a starting material for the production of many commercially important nitrogen compounds. Since the
improvement potentials of renewable technologies are most likely greater than fossil fuels, it is significant to implement
renewable-based alternative fuel production options from the environmental point of view. In this case study, maritime vehicle,
freight ship, is driven with ammonia instead of heavy fuel oils in the power engines. Additionally, dual-fuel options, heavy fuel oil,
and ammonia are investigated. A comparative LCA of transoceanic freight ship is performed to examine the effects of clean fuel
driven maritime vehicles on the environment. The complete transport life cycle is evaluated in the life cycle analyses comprising
manufacture of freight ship; operation of freight ship; construction and land use of seaport; operation, maintenance, and disposal
of a seaport; and production and transportation of ammonia. Ammonia is produced using renewable resources, namely, biomass,
hydropower, municipal waste and geothermal.

2.1.4.2 Pathways for Production of Ammonia


Currently, SMR is a more common method for ammonia production; developing electrolyzers can be used for water electrolysis
driven by renewable energy resources. Renewable resources have lower environmental impacts. Therefore, municipal waste,
biomass, hydropower and geothermal sources are utilized for ammonia production. For hydrogen production, electrolysis route is
employed where an electrolyzer is used with an energy requirement of 53 kWh electricity to generate 1 kg of hydrogen. The source
of electricity is taken from municipal waste, geothermal, biomass and hydropower individually for all cases.
The chief commercial method of producing ammonia is by the Haber–Bosch process, which involves the direct reaction of
elemental hydrogen and elemental nitrogen.

N2 þ 3H2 -2NH3 ð31Þ

This reaction requires the use of a catalyst, high pressure about 200 atm and elevated temperature about 4501C. Generally, the
catalyst is iron containing iron oxide. The Haber–Bosch process is utilized for ammonia synthesis in this case study. Hydrogen is
taken from electrolysis unit and nitrogen is supplied through the air separation process. Cryogenic air separation is typically
chosen technique for a huge quantity of nitrogen manufacture. For the LCA of nitrogen manufacture, electrical work for the
procedure, cooling water, surplus heat, and groundwork for the air separation facility are taken into account. The distribution
elements were attained from the heat of vaporization and the specific heat capacity multiplied with the temperature difference
from 201C to the boiling point [76]. Cryogenic air separation process becomes more cost effective compared to non-cryogenic
methods at the level of about 200–300 t per day nitrogen. Since gas phase nitrogen is required in the ammonia production
30 Ammonia

reaction, the energy requirement is lower compared to liquid nitrogen because liquefaction is not required. The major input to the
air separation plant is electricity required to compress the air. Here, US mix grid electricity is used for air separation plant. Air is not
taken as an input because of inexhaustibility. The separated CO2 and water vapor are not evaluated as emissions in the process.
The transportation of nitrogen is not accounted for the analyses since it is considered that the ammonia synthesis plant is located
near air separation plant. The transportation of ammonia to the port are also considered in the analyses. 0.6 tonne-kilometer
(tkm) via diesel driven rail transport and 0.1 tkm lorry transport (higher than 16 t) per kg of ammonia are considered. Producing
liquid products from air separation plant requires about two times higher energy than gaseous products. Commercial cryogenic air
separation plants require electricity in the range 0.6–1 kWh kg1 of liquid nitrogen product. However, as mentioned earlier,
gaseous product necessitates lower power input. Hence, in this case study, 0.42 kWh electricity is assumed for nitrogen gas
production as taken from the GREET 2016 model [83]. The utilized electricity for nitrogen production is US mix grid. For
ammonia generation, municipal waste, geothermal, hydropower, and biomass power plant electricity is used from US power
plants. For the transportation of the produced ammonia, an average distance is assumed where 100 km is by lorry with a capacity
of higher than 16 t and 600 km by rail transport. Dual fuel operation of vessels is also considered in the study as 50% ammonia
and 50% heavy fuel oil.

2.1.4.3 Material and Methods


LCA is a powerful method to inspect environmental impacts of a system or process or product. LCA represents a methodical set of
processes for assembling and examining the inputs and outputs of materials and energy, and the related environmental impacts,
directly assignable to the product or service during the course of its life cycle. A life cycle is the set of phases of a product or service
system, from the extraction of natural resources to last removal. Entire life cycle steps from resource extraction to disposal during
the lifetime of a product or process are considered in this case study. LCA is a four-step process, namely, goal and scope definition,
inventory analysis, impact assessment, improvement potential.
The goal of this case study is to explore the environmental effects of ammonia fueled marine transportation ships in
comparison with conventional heavy fuel oil from cradle to grave using GWP, abiotic depletion, acidification, stratospheric
ozone layer depletion, marine eco-toxicity, and marine sediment ecotoxicity. The motivation behind this case study is to decrease
the environmental impacts caused by current hydrocarbon dependent marine transportation systems. The results of this case
study will mainly attract marine transportation sector and academicians working in the area of clean fuel production and
utilization technologies. The function of this case study assesses environmental impacts per tonne-kilometer cruise travel where the
functional unit is 1 tkm. It is assumed that the processes for ammonia production contain production of hydrogen and nitrogen
separately and the mass balance is used to identify the amount of hydrogen and nitrogen required for 1 kg of ammonia
production.
For this analysis, we account for all the stages in the life cycle of maritime transportation, including feedstock recovery and
transportation, fuel production and transportation, and fuel consumption in the ocean freight ships. The exploration and recovery
activities from the well to fuel production and the subsequent transportation to the pump constitute the WTP stage. The com-
bustion of fuel during ocean vehicle operation constitutes the pump-to-haul (PTH) stage. These two stages combined comprise the
well-to-haul (WTH) cycle. There are numerous assessment procedures established over the time to categorize and depict the
environmental effects of the system. The method used for the present analysis is CML 2001 which is a method proposed by a set of
scientists under the principal of CML (Center of Environmental Science of Leiden University).
The system boundaries show the limitations of unit processes which are needed to be included in an LCA study. The system
boundaries for the LCA analyses are defined as shown in Fig. 28, namely: mining of the raw materials and extraction of the
nutrients from these materials, transportation of raw materials and pre-products, generation, and supply of required energy,
manufacturing of the ammonia and the related field operations.
Various environmental impact categories including global warming, marine sediment ecotoxicity, marine aquatic ecotoxicity,
acidification, ozone layer depletion, and abiotic depletion are selected in order to study the diverse effects of changing to ammonia
fuels in sea transportation. The description of the categories is already explained in case study 1.

Ship maintenance Fuel transport Port construction


Port
maintenance

NH3
Ship
operation
Heavy fuel
oil

Port
operation
Ship
manufacturing Fuel production Port disposal

Fig. 28 Life cycle steps of maritime transportation.


Ammonia 31

Table 7 Distance, speed, and duration of trip for energy requirement of transoceanic freight ship

Distance (nmi) Speed (kn) Time (h) Load factor

Cruise 2306.041 18.006 128.073 0.6


RSZ1 100.827 18.006 5.6 0.6
RSZ2 25 18.006 1.388 0.6
Hotel1 – – 22 0.19
Hotel2 – – 22 0.19

Source: Data from GREET 2016. Argonne, IL: Argonne National Laboratory; 2016.

2.1.4.4 Life Cycle Assessment Phases


Analysis of WTH of ammonia fuels is performed in this case study where each step is briefly explained in this section.
The functional unit in the LCA study is 1 tkm marine transportation. A ton kilometer by shipping is defined as unit of measure of
goods transport which represent the transport of 1 t by a vessel over 1 km. Marine diesel engines are generally further categorized
into two different groups as slow speed (15 knots in average) and medium speed (25–30 knots). Transoceanic freight ships are
slow speed vehicles which include two-stroke cycle with crosshead engines of 4–12 cylinders. In the marine industry, these engines
are used for main propulsion and constitute a larger portion of installed power on the ship.
Each vessel type is characterized by the power rating of its two engines – main and auxiliary. For each engine, fuel consumption
and emission factors are defined. The GREET 2016 software is utilized to find the power ratings and total energy consumptions of
the selected ships [83].
In order to determine the average power consumptions, the trip is assumed to be from Pacific to international ports. After
determining the total trip distances as listed in Table 7, each trip was divided into segments, including transit through reduced-
speed zones (RSZs). Each trip segment may have distinct fuel consumption and emission factors owing to different speeds and
load factors and engine/fuel switching. At the origin and destination ports, the vessel will leave hotel and burn fuel dockside using
mainly auxiliary engines. After a ship leaves port, it travels in an RSZ, during which it uses a lower load factor and consumes less
fuel, thereby emitting fewer pollutants, than when traveling at cruising speed. It will pass through an RSZ before hoteling at the
port of destination. The GREET marine module aims to model these trip segments, representatively, for different vessel types
leaving or arriving at U.S. ports in the various regions [83].

2.1.4.4.1 Transoceanic freight ship manufacture


Transoceanic freight ship manufacture includes the processes of material production, representing the material composition of
an average water vehicle used for solid goods transportation. For manufacturing, electricity and heavy oil burned in the
industrial furnace are included. For the transportation of materials, standard distances are applied. Also, waste treatment
processes for non-metal components of a water vehicle are accounted for. The exchanges of the ship manufacturing are
derived from an assessment of a ship, with a load capacity of 51,500 t. The energy consumption in manufacturing is
estimated as 50% of the cumulative energy of the used materials. The split of energies is 10% electricity and 90% heavy fuel
oil [76].

2.1.4.4.2 Maintenance of transoceanic freight ship


Maintenance of transoceanic freight ship includes the use paint and emissions of the solvent of the paint as NMVOC. Con-
sumption of lubricates is excluded. It is assumed that the ship is painted six times in its entire lifespan of 25 years. The
consumption of lubricates is included in the fuel consumption for vessel operation [76].

2.1.4.4.3 Port facilities


Port facilities comprise the construction and disposal of one of the world’s biggest port in Rotterdam, Netherlands. The inventory
contains the processes of material production, representing the material used in the construction phase of the harbor. The building
activities and electricity consumption are accounted for building and disposal phases. Emission of NMVOC is included. Also, lorry
transport of materials to the constructions site are taken into account. The expenditures due to construction and disposal are
addressed. The data represent one seaport for both, sea and inland shipping. The material composition of the sealed industrial area
is derived from the construction and disposal expenditures of a European highway. The built-up is modeled as a steel building. The
lifetime of port is assumed to be 100 years [76].

2.1.4.4.4 Maintenance and operation of the port


Maintenance and operation of the port include emissions to water due to non-removed oil spills. In this step, the land occupation
and transformation due to seaport are taken into account. The energy consumption at the port is based on assumptions for the
specific electricity consumption at the harbor of Hamburg, Germany. The land use is further distinguished in a built-up area
(0.6%), road area (45.6%) and water bodies (54.29%). Emission to waters includes emission from production sites on the port
site, which are not directly connected to the transport activities [76].
32 Ammonia

2.1.4.4.5 Operation of the transoceanic freight ship and transport for 1 tkm
In this phase, the full cycle is represented by the supply of the fuel, the operation of the ship and transport of the goods as one tkm.
Direct airborne emissions of gaseous substances, particulate matters, dioxins, PAHs, halogens and heavy metals are accounted for.
These emissions are caused by heavy oil burning in the engines. Hence they are mostly eliminated in ammonia driven ships. Also,
the disposal of bilge oil and emissions of tributyltin compounds are included. Individual hydrocarbons are estimated based on the
share of diesel engines of road vehicles. Heavy metals are estimated from trace elements in fuel. A distinction between distilled
(28%) and residual fuel (72%) is applied. The amount of disposed bilge oil is estimated as 0.6% of the consumed fuel. The
average data for the steam turbine (5%) and diesel engine (95%) propulsion are considered in the study. The fuel used for
conventional ships is heavy fuel oil and is representative of slow speed engine types about 15 knots. The data represents solid bulk
transport of about 40,000 dwt (deadweight tonnage) where the ship is driven by the steam turbine and diesel engines [76]. The
power ratings of the main engine and auxiliary engines are about 37.5 and 8.3 MW, respectively. The average energy consumption
for the freight ship is calculated to be 0.214 MJ per mile ton [83]. The entire transport life cycle namely; the operation of the vessel;
production of the vessel; construction and land use of port; operation, maintenance, and disposal of the port; production and
transportation of fuel to the port are considered. Port infrastructure expenditures and environmental interventions are allocated
based the yearly throughput (0.37). The vessel manufacturing is allocated based on the total kilometric performance corre-
sponding to about 2,000,000 km and its transport performance. Since transport activity requires loading and unloading, for each
transport activity two ports are needed [76].

2.1.4.5 Case Study 2 Results and Discussion


The life cycle analyses are performed in the SimaPro and GREET 2016 software for the assessment of environmental impacts. Power
consumption and emission values for transoceanic freight ships are derived from GREET 2016 based on the trip scenario listed in
Table 7. The obtained values are used in the SimaPro LCA software by employing the impact assessment method of CML 2001.
The toxic substances on the marine sediment and aquatic environment are the main concerns of marine sediment ecotoxicity
and marine aquatic ecotoxicity categories. 1,4-Dichlorobenzene equivalents/tkm is used to express each toxic substance. Among
the selected fuels, the conventional heavy fuel oil for freight ship have the greatest damage on marine sediment and aquatic
environment as shown in Figs. 29–34.
Ammonia driven vehicles where ammonia is produced from geothermal and municipal waste yielded the lowest marine
sediment ecotoxicity impact with a value of about 0.0047 kg 1,4-DB eq. tkm1 for transoceanic freight ship. Using ammonia where
it is produced from the municipal waste plant, in the transoceanic ship as dual fuel with heavy fuel oil lowers the marine sediment
ecotoxicity level about 49%.
The contributions of different processes to ecotoxicity of marine sediment and marine aqua are shown in Figs. 30–32. The
operation of the freight ship is responsible for 45% of marine aquatic ecotoxicity as seen in Fig. 32 for sole ammonia fueled ship
whereas exploration and offshore production of heavy oil represents 6% and natural gas extraction represents 15% of total. This is
due to natural gas, and oil-fired power plants which are then used for nitrogen production plant. For biomass-based ammonia
driven freight ship, barium, tributyltin compounds, and vanadium are top three substances causing marine aquatic ecotoxicity as
where barium has an impact corresponding to 0.0017 kg 1,4-DB eq. tkm1. Some of them are related to tributyltin compounds
emitted to water because of bottom paintings of the ships.
Fig. 30 comparatively shows marine sediment ecotoxicity values of subprocesses for ship powered by sole ammonia from
municipal waste whereas Fig. 31 illustrates for dual fuel option (from hydropower). The operation of freight ship, in this case,

Conventional heavy fuel oil

Ammonia (hydropower)/heavy fuel oil

Ammonia (hydropower)

Ammonia (municipal waste/heavy fuel oil dual fuel)

Ammonia (municipal waste)

Ammonia (geothermal/heavy fuel oil dual fuel)

Ammonia (geothermal)

Ammonia (biomass/heavy fuel oil dual fuel)

Ammonia (biomass)
0

25

75

01

25

75

02
00

01
00

00

0.

01

01

0.
0.

0.
0.

0.

0.

0.

Marine sediment ecotox. 500a (kg 1,4-DB eq. tkm–1)

Fig. 29 Marine sediment ecotoxicity values of transoceanic freight ship per ton kilometer (tkm) for ammonia and conventional heavy fuel oil.
Ammonia 33

Operation,
0.003 transoceanic freight
ship, 0.0028

Marine sediment ecotoxicity (kg 1,4-DB eq. tkm–1)


0.0025

0.002

0.0015
Remaining
processes, 0.0012

0.001
Natural gas,
unprocessed, at
extraction, 0.0005 Heavy fuel oil,
0.0005 burned in power
plant, 0.0003

0
Fig. 30 Process contributions to marine sediment ecotoxicity values of transoceanic freight ship fueled by sole ammonia from municipal waste.

Operation, transoceanic freight ship

Well for exploration and production,


offshore
10%
Natural gas, unprocessed, at extraction
3% 4%
5% Heavy fuel oil, burned in power plant
2%
1%
6% Discharge, produced water, onshore

Discharge, produced water, offshore


1%
73%
1% Heavy fuel oil, burned in refinery
furnace
Heavy fuel oil, burned in industrial
furnace 1MW
Remaining processes
Fig. 31 Process contributions to marine sediment ecotoxicity values of transoceanic freight ship driven by dual fuel (50% ammonia from
hydropower and 50% heavy fuel oil).

causes more than 80% of total ecotoxicity. The remaining contributor is mainly operation/maintenance of the port corresponding
to about 14% because of heavy fuel oil and natural gas-fired power plants for the electricity requirement of the port. The operation
of the freight ship is responsible for 73% of marine sediment ecotoxicity for ammonia/heavy fuel oil driven ship whereas
exploration, and offshore production of heavy oil represents the 10%. This is due to crude oil production which is then used in a
heavy fuel oil refinery and transported to heavy fuel oil regional storage to be combusted in the ship.
Ecotoxicity level of marine aqua is significantly lower when ammonia from municipal waste, hydropower and geothermal are
utilized in the engines as seen in Fig. 32. This validates that using ammonia as a supplementary fuel to heavy oil decreases the total
environmental impact significantly. The process contributions to marine aquatic ecotoxicity values of transoceanic freight ship
fueled by sole ammonia from geothermal energy are shown in Fig. 33.
Regarding GWP, ammonia (from geothermal energy) fueled freight ship yield the lowest greenhouse gas emissions in the entire
life cycle corresponding to 0.0043 kg CO2 eq. tkm1 for freight ship as shown in Fig. 34. However, it is very high for the
conventional heavy fuel oil. The highest values after sole heavy fuel oil are 0.0094 kg CO2 eq. tkm1 for ammonia from biomass
and heavy fuel oil combination, respectively, for freight ship.
The acidification values of heavy fuel oil driven transoceanic freight ship are mainly caused by SO2 and NOx emissions which
correspond to more than 90% of overall acidification value where comparative results are shown in Fig. 35. The source of SO2
34 Ammonia

Conventional heavy fuel oil


Ammonia (hydropower)/heavy fuel oil
Ammonia (hydropower)
Ammonia (municipal waste/heavy fuel oil dual fuel)
Ammonia (municipal waste)
Ammonia (geothermal/heavy fuel oil dual fuel)
Ammonia (geothermal)
Ammonia (biomass/heavy fuel oil dual fuel)
Ammonia (biomass)

8
2

01
0

00

00

01

01

01

01
00

00

0.
0.

0.

0.

0.

0.

0.
0.

0.
Marine aquatic ecotox. 500a
(kg 1,4-DB eq. t km–1)
Fig. 32 Marine aquatic ecotoxicity values of transoceanic freight ship per ton kilometer (tkm) for ammonia and conventional heavy fuel oil.

Remaining Operation,
processes transoceanic freight
27% ship
45%

Well for exploration


and production,
offshore
6%

Heavy fuel oil,


burned in power
plant
7%
Natural gas,
unprocessed, at
extraction
15%

Fig. 33 Process contributions to marine aquatic ecotoxicity values of transoceanic freight ship fueled by sole ammonia from geothermal energy.

Conventional heavy fuel oil


Ammonia (hydropower)/heavy fuel oil
Ammonia (hydropower)
Ammonia (municipal waste/heavy fuel oil dual fuel)
Ammonia (municipal waste)
Ammonia (geothermal/heavy fuel oil dual fuel)
Ammonia (geothermal)
Ammonia (biomass/heavy fuel oil dual fuel)
Ammonia (biomass)

0 0.002 0.004 0.006 0.008 0.01 0.012


–1
Global warming 500a (kg CO2 eq. tkm )

Fig. 34 Global warming potential (GWP) of transoceanic freight ship per ton kilometer (tkm) for ammonia and conventional heavy fuel oil.
Ammonia 35

Conventional heavy fuel oil


Ammonia (hydropower)/heavy fuel oil
Ammonia (hydropower)
Ammonia (municipal waste/heavy fuel oil dual fuel)
Ammonia (municipal waste)
Ammonia (geothermal/heavy fuel oil dual fuel)
Ammonia (geothermal)
Ammonia (biomass/heavy fuel oil dual fuel)
Ammonia (biomass)
0 0.00008 0.00016 0.00024
–1
Acidification (kg SO2 eq. tkm )

Fig. 35 Acidification values of transoceanic freight ship per ton kilometer (tkm) for ammonia and conventional heavy fuel oil.

Conventional heavy fuel oil

Ammonia (hydropower)/heavy fuel oil

Ammonia (hydropower)

Ammonia (municipal waste/heavy fuel oil dual fuel)

Ammonia (municipal waste)

Ammonia (geothermal/heavy fuel oil dual fuel)

Ammonia (geothermal)

Ammonia (biomass/heavy fuel oil dual fuel)

Ammonia (biomass)

0.00E+00 3.00E−05 6.00E−05 9.00E−05


–1
Abiotic depletion (kg Sb eq. tkm )

Fig. 36 Abiotic depletion values of transoceanic freight ship per ton kilometer (tkm) for ammonia and conventional heavy fuel oil.

emission is predominantly the operation of freight ship (96.8%) for conventional heavy fuel oil ships. This is caused by the sulfur
content of the heavy fuel oil hence it is mostly eliminated in clean fuels as seen in Fig. 35. The combustion of diesel and heavy fuel
oil have the high impact hence, particularly transportation processes with railway and trucks create high emissions leading to
higher acidification values.
The abiotic sources are natural sources counting energy sources, such as hard coal and crude oil, which are evaluated as non-
living. This is because of fossil fuels are major basis of energy and feed resource, it shows the huge intake of hard coal and lignite
for tonne-kilometer travel of ammonia (from biomass) fueled transoceanic freight ship as shown in Fig. 36.
The reason of hard coal and lignite consumption is the electricity US mix usage in air separation plant for nitrogen production
of ammonia synthesis process. Heavy fuel oil utilization constitutes about 53% of overall abiotic depletion whereas ammonia
production from biomass power plant electricity constitutes about 12.3% of total abiotic depletion for ammonia (from biomass)/
heavy fuel oil combination. Also, operation and maintenance of the port have also a high share corresponding to 31.5% of total
where it is similarly originated from electricity mix production.
Fig. 37 presents the life cycle kg CFC-11 eq. emissions of the freight ship with ammonia and conventional fuel oil per tonne-
kilometer traveled. It is quite high for heavy fuel oil and dual-fuel options while it is considerably less for ammonia fuel.
Particularly, hydropower options have the lowest environmental impact in terms of ozone layer depletion.
For ammonia and heavy fuel oil driven transoceanic freight ship, crude oil production has the highest share corresponding to
about 81.2% as shown in Fig. 38. The operation and maintenance of the port are responsible for about 16.6% of overall ozone
layer depletion whereas the manufacturing of the ship constitutes only 2.2%. Two main substances causing ozone layer depletion
are bromotrifluoromethane (Halon 1301) and bromochlorodifluoromethane (Halon 1211) corresponding to about 1.98E  10
and 3.31E  11 kg CFC-11 eq. tkm1, respectively.

2.1.4.6 Case Study 2 Conclusions


Reducing the total greenhouse gas emissions from marine transportation is possible using ammonia which is carbon-free fuel.
They can be utilized for maritime ship engines directly as supplementary fuels or individual fuels. Ammonia fueled ships yield
considerably lower global warming impact during operation. The highest GWPs are calculated to be 0.010 kg CO2 eq. tkm1 for
conventional heavy fuel oil, respectively, for transoceanic freight ship. Using ammonia as a dual fuel in the marine engines can
36 Ammonia

Conventional heavy fuel oil


Ammonia (hydropower)/heavy fuel oil
Ammonia (hydropower)
Ammonia (municipal waste/heavy fuel oil dual fuel)
Ammonia (municipal waste)
Ammonia (geothermal/heavy fuel oil dual fuel)
Ammonia (geothermal)
Ammonia (biomass/heavy fuel oil dual fuel)
Ammonia (biomass)

0.00E+00 4.00E−10 8.00E−10 1.20E−09 1.60E−09

Ozone layer depletion 40a (kg CFC-11 eq. tkm–1)


Fig. 37 Stratospheric ozone layer depletion values of transoceanic freight ship per ton kilometer (tkm) for ammonia and conventional heavy fuel oil.

4.0E−10
Crude oil, at
production onshore
Ozone layer depletion (kg CFC-11 eq. tkm–1)

3.5E−10

3.0E−10

2.5E−10

2.0E−10

1.5E−10

1.0E−10
Transport, natural
gas, pipeline, long Crude oil, at Remaining
5.0E−11 distance processes
production

0.0E+00
Fig. 38 Process contributions to stratospheric ozone layer depletion of transoceanic freight ship by dual fuel (50% ammonia from municipal
waste and 50% heavy fuel oil).

decrease total greenhouse gas emissions up to 33.5% (for geothermal-based ammonia) whereas this number increases to 69% if
only ammonia (from geothermal) is used in the engines. The LCA study proves that switching to ammonia in maritime trans-
portation reduces the total GHG emissions and other environmental impacts considerably.

2.1.5 Closing Remarks

Ammonia becomes a primary hydrogen carrier that does not contain any carbon atoms and has a high hydrogen ratio. It consists
of one nitrogen atom from air separation and three hydrogen atoms from any conventional or renewable resources. Ammonia as a
sustainable fuel can be used in all types of combustion engines, gas turbines, burners with only small modifications and directly in
fuel cells which is a very significant advantage compared to another type of fuels. It is also an option for cooling the engine with
ammonia that can act as a refrigerant while it is heated to the temperature at which it is fed to the power producing unit in the
vehicle. Reducing the total greenhouse gas emissions from marine transportation is possible using ammonia which is carbon-free
fuel. They can be utilized for maritime ship engines directly as supplementary fuels or individual fuels. Ammonia fueled ships
yield considerably lower global warming impact during operation. Ammonia as a sustainable and clean fuel in road vehicles yield
also the lowest GWP after electric and hydrogen vehicles. Moreover, the overall thermal and exergy efficiencies of the ion and
proton-conducting direct ammonia SOFC range of 70–85% depending on the operating conditions indicating the suitable
arrangement of components of the systems.
Ammonia 37

Acknowledgment

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

References

[1] Veziroglu TN, Şahin S. 21st Century’s energy: Hydrogen energy system. Energy Convers Manag 2008;49:1820–31.
[2] Jensen JO, Vestbø AP, Li Q, Bjerrum NJ. The energy efficiency of onboard hydrogen storage. J Alloys Compd 2007;446:723–8.
[3] Zamfirescu C, Dincer I. Ammonia as a green fuel and hydrogen source for vehicular applications. Fuel Process Technol 2009;90:729–37.
[4] Gálvez ME, Halmann M, Steinfeld A. Ammonia production via a two-step Al2O3/AlN thermochemical cycle. 1. Thermodynamic, environmental, and economic analyses. Ind
Eng Chem Res 2007;46:2042–6.
[5] Kool A, Marinussen M, Blonk H. LCI data for the calculation tool Feedprint for greenhouse gas emissions of feed production and utilization. GHG Emissions of N, P and
K Fertilizer Production Blank Consultants Gravin Beatrixstraat 2012;34.
[6] Ganley JC, Holbrook JH, McKinley DE. 2007. Solid state ammonia synthesis. In: Ammonia fuel network conference; 2007.
[7] Gilbert P, Thornley P. Energy and carbon balance of ammonia production from biomass gasification. In: Bio-ten; 21 Sep 2010-23 Sep 2010; Birmingham: Holiday Inn;
2010.
[8] Anderson K, Bows A, Mander S. From long-term targets to cumulative emission pathways: reframing UK climate policy. Energy Policy 2008;36:3714–22.
[9] Thomas G, Parks G. Potential roles of ammonia in a hydrogen economy. Washington, DC: US Department of Energy; 2006.
[10] Lan R, Tao S. Ammonia as a suitable fuel for fuel cells. Front Energy Res 2014;2:35.
[11] Hacker V, Kordesch K. Ammonia crackers. In: Vielstich W, Gasteiger HA, Lamm A, editors. Handbook of fuel cells. Chichester: John Wiley & Sons, Ltd.; 2010.
doi:10.1002/9780470974001.f302011.
[12] Home – Industry Overview – PotashCorp. Available From: http://www.potashcorp.com/overview/introduction/company/potashcorp-profile; 2017 [accessed 07.01.17].
[13] Bomelburg HJ. Use of ammonia in energy-related applications. Plant/Operations Prog 1982;1:175–80.
[14] Schlapbach L, Züttel A. Hydrogen-storage materials for mobile applications. Nature 2001;414:353–8.
[15] Feibelman PJ, Stumpf R. Comments on potential roles of ammonia in a hydrogen economy – a study of issues related to the use of ammonia for on-board vehicular
hydrogen storage. Sandia National Laboratories; 2006.
[16] Christensen CH, Johannessen T, Sørensen RZ, Nørskov JK. Towards an ammonia-mediated hydrogen economy? Catal Today 2006;111:140–4.
[17] Blarigan P Van. Advance internal combustion engine research. In: Proc. 2000 DOE Hydrogen Program Review; 2000. p. 1–19.
[18] McFarlan A. Development of direct ammonia fuel cells for efficient stationary CHP applications. Proceedings of the ammonia–sustainable, emission free fuel conference,
October. San francisco, CA; 2007.
[19] Olson N. BioAmmonia–a comparison with other biofuels. In: Proc. 2007 annual NH3 fuel conference. Iowa Energy Center, San Francisco; 2007.
[20] Lide DR. CRC handbook of chemistry and physics: a ready-reference book of chemical and phyical data. Boca Raton, FL: CRC Press; 1990.
[21] Elmøe TD, Sørensen RZ, Quaade U, et al. A high-density ammonia storage/delivery system based on Mg(NH3)6Cl2 for SCR–DeNOx in vehicles. Chem Eng Sci
2006;61:2618–25.
[22] Zhu L, Wang S, Gu J. Performance investigation of a thermal‐driven refrigeration system. Int J Energy Res 2008;32:939–49.
[23] Park YM, Sonntag RE. A preliminary study of the kalina power cycle in connection with a combined cycle system. Int J Energy Res 1990;14:153–62.
[24] Hilali İ, Söylemez MS. On the optimum sizing of exhaust gas-driven automotive absorption cooling systems. Int. J. Energy Res 2008;32:655–60.
[25] Zamfirescu C, Dincer I. Environmentally-benign hydrogen production from ammonia for vehicles. In: Proc. global conference on global warming; 2008. p. 6–10.
[26] Zamfirescu C, Dincer I. Using ammonia as a sustainable fuel. J Power Sources 2008;185:459–65.
[27] Zamfirescu C, Dincer I. Ammonia as a green fuel for transportation. In: ASME 2008 2nd Int. conference on energy sustainability, vol. 1, ASME; 2008. p. 507–15.
doi:10.1115/ES2008-54328.
[28] Williams OM, Carden PO. Energy storage efficiency for the ammonia/hydrogen-nitrogen thermochemical energy transfer system. Int J Energy Res 1979;3:29–40.
[29] Williams OM, Carden PO. Ammonia dissociation for solar thermochemical absorbers. Int J Energy Res 1979;3:129–42.
[30] Christensen CH, Sorensen RZ, Johannessen T, et al. Metal ammine complexes for hydrogen storage. J Mater Chem 2005;15:4106–8.
[31] Appl M. Ammonia: principles and industrial practice. Weinheim: Vch Verlagsgesellschaft Mbh; 1999.
[32] Strickland GG. 1980. Ammonia as a hydrogen energy-storage medium. Present. 5th annual thermal storage meeting, McLean, VA; 1980.
[33] Metkemeijer R, Achard P. Ammonia as a feedstock for a hydrogen fuel cell; reformer and fuel cell behaviour. J. Power Sources 1994;49:271–82.
[34] Waitzman DA. Ammonia from coal: a technical/economic review. Process technology and flow sheets. In: Cavaseno V, editor. Chemical engineering. New York: McGraw-
Hill Publ. Co; 1979. p. 63–5.
[35] Skodras G, Kaldis S, Topis S, et al. NH3 decomposition and simultaneous H2 separation with a commercial Pd–Cu–Ag/V membrane. In: Proc. Second Int. Green Energy
Conference; 2006. p. 25–9.
[36] Vitse F, Cooper M, Botte GG. On the use of ammonia electrolysis for hydrogen production. J Power Sources 2005;142:18–26.
[37] Canadian Industry Program for Energy Conservation. Benchmarking energy efficiency and carbon dioxide emissions, Canada; 2008.
[38] Ammonia|Industrial Efficiency Technology Database; Measures. Available From: http://ietd.iipnetwork.org/content/ammonia#key-data; 2017 [accessed 7.01.17].
[39] Olson NK, Holbrook J. NH3 – the other hydrogen. Richland, WA: Ammonia Fuel Network; 2008.
[40] Dincer I, Zamfirescu C., editors. 2011. Ammonia as a potential substance. In: Sustainable energy systems and applications. Boston, MA: Springer US; p. 203–32.
doi:10.1007/978-0-387-95861-3_7.
[41] Heldebrant DJ, Karkamkar A, Linehan JC, Autrey T. Synthesis of ammonia borane for hydrogen storage applications. Energy Environ Sci 2008;1:156–60.
[42] PotashCorp Integrated Annual Report. Annual Integrated Report. Available From: http://www.potashcorp.com/irc/nitrogen; 2015 [accessed 07.01.17].
[43] International Energy Agency. Energy Technology Perspectives 2012; Pathways to a Clean Energy System, France. ISBN: 978-92-64-17488-7.
[44] Dossat RJ, Horan TJ. Principles of refrigeration. Upper Saddle River, NJ: Prentice Hall; 2002.
[45] Yin SF, Xu BQ, Zhou XP, Au CT. A mini-review on ammonia decomposition catalysts for on-site generation of hydrogen for fuel cell applications. Appl Catal A Gen
2004;277:1–9.
[46] Klerke A, Christensen CH, Norskov JK, Vegge T. Ammonia for hydrogen storage: challenges and opportunities. J Mater Chem 2008;18:2304–10.
[47] Sakintuna B, Lamari-Darkrim F, Hirscher M. Metal hydride materials for solid hydrogen storage: a review. Int J Hydrogen Energy 2007;32:1121–40.
[48] Marder TB. Will we soon be fueling our automobiles with ammonia–borane? Angew Chemie Int Ed 2007;46:8116–8.
[49] Best Available Techniques for Pollution Prevention and Control in the European Fertilizer Industry PRODUCTION OF AMMONIA. Brussels, Belgium. 2000.
[50] Ammonia: OSH Answers. Can. Cent. Occup. Heal. Saf. Available From: http://www.ccohs.ca/oshanswers/chemicals/chem_profiles/ammonia.html; 2017
[accessed 07.01.17].
[51] Putnam DF. Composition and concentrative properties of human urine Report NASA CR-1802. Washington, DC: National aeronautics and space administration; 1971.
38 Ammonia

[52] Ishak F. Thermodynamic analysis of ammonia and urea fed solid oxide fuel cells; Master of applied science (MASc) dissertation, University of Ontario Institute of
Technology; 2011.
[53] EG&G Technical Services I. Fuel cell handbook. Morgantown, West Virginia: US Department of Energy; 2004.
[54] Barbir F. PEM fuel cells: theory and practice. London: Academic Press; 2013.
[55] Mench MM. Fuel cell engines. Hoboken, NJ: John Wiley & Sons; 2008.
[56] Ni M, Leung DYC, Leung MKH. Electrochemical modeling and parametric study of methane fed solid oxide fuel cells. Energy Convers Manag 2009;50:268–78.
[57] Arpornwichanop A, Patcharavorachot Y, Assabumrungrat S. Analysis of a proton-conducting SOFC with direct internal reforming. Chem Eng Sci 2010;65:581–9.
[58] Chan S, Low C, Ding O. Energy and exergy analysis of simple solid-oxide fuel-cell power systems. J Power Sources 2002;103:188–200.
[59] Singhal SC, Kendall K. High-temperature solid oxide fuel cells: fundamentals, design, and applicatons. Oxford, UK: Elsevier Advanced Technology; 2003.
[60] Suwanwarangkul R, Croiset E, Fowler MW, et al. Performance comparison of Fick’s, dusty-gas and Stefan–Maxwell models to predict the concentration overpotential of a
SOFC anode. J Power Sources 2003;122:9–18.
[61] Ho CK, Webb SW, Stephen W. Gas transport in porous media. Berlin: Springer; 2006.
[62] Janardhanan V. A detailed approach to model transport, heterogeneous chemistry, and electrochemistry in solid-oxide fuel cells. Karlsruhe: Universitätsverlag; 2007.
[63] Bird RB, Robert B, Stewart WE, Lightfoot EN. Transport phenomena. New York: J. Wiley; 2007.
[64] Wu M, Wang M, Liu J, Huo H. Assessment of potential life-cycle energy and greenhouse gas emission effects from using corn-based butanol as a transportation fuel.
Biotechnol Prog 2008;24:1204–14.
[65] Poling Bruce E, Prausnitz John M, O’Connell John P. Properties of gases and liquids. New York: McGraw-Hill Education; 2001.
[66] Zhu H, Kee RJ, Janardhanan VM, Deutschmann O, Goodwin DG. Modeling elementary heterogeneous chemistry and electrochemistry in solid-oxide fuel cells. J
Electrochem Soc 2005;152:A2427.
[67] Li M, Zhang X, Li G. A comparative assessment of battery and fuel cell electric vehicles using a well-to-wheel analysis. Energy 2016;94:693–704.
[68] Zhu H, Kee RJ. A general mathematical model for analyzing the performance of fuel-cell membrane-electrode assemblies. J Power Sources 2003;117:61–74.
[69] Ni M, Leung DYC, Leung MKH. An improved electrochemical model for the NH3 fed proton conducting solid oxide fuel cells at intermediate temperatures. J Power
Sources 2008;185:233–40.
[70] Patcharavorachot Y, Paengjuntuek W, Assabumrungrat S, Arpornwichanop A. Performance evaluation of combined solid oxide fuel cells with different electrolytes. Int J
Hydrogen Energy 2010;35:4301–10.
[71] Demin A, Tsiakaras P. Thermodynamic analysis of a hydrogen fed solid oxide fuel cell based on a proton conductor. Int J Hydrogen Energy 2001;26:1103–8.
[72] Chein R-Y, Chen Y-C, Chang C-S, Chung JN. Numerical modeling of hydrogen production from ammonia decomposition for fuel cell applications. Int J Hydrogen Energy
2010;35:589–97.
[73] Rose L, Hussain M, Ahmed S, et al. A comparative life cycle assessment of diesel and compressed natural gas powered refuse collection vehicles in a Canadian city.
Energy Policy 2013;52:453–61.
[74] GREET 2015. Argonne, IL: Argonne National Laboratory; 9700 S. Cass Avenue, Building 362, Argonne, IL 60439-4844 USA 2015.
[75] Archsmith J, Kendall A, Rapson D. From cradle to junkyard: assessing the life cycle greenhouse gas benefits of electric vehicles. Res Transp Econ 2015;52:72–90.
[76] Consultants P. SimaPro life cycle analysis version 7.2 (software). PRé Sustainability, Stationsplein 121, 3818 LE Amersfoort, The Netherlands.
[77] 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].
[78] ISO 14044:2006; Environmental management – Life cycle assessment – Requirements and guidelines. Available From: http://www.iso.org/iso/catalogue_detail?
csnumber=38498; 2006 [accessed 07.01.17].
[79] Guinée JB, Heijungs R, Huppes G, et al. Life cycle assessment: past, present, and future. Environ Sci Technol 2011;45:90–6.
[80] Duce AD, Egede, P, Öhlschläger G, et al. eLCAr – Guidelines for the LCA of electric vehicles. Report from Project. E-Mobility Life Cycle Assessment Recommendations
funded within European Union Seventh Framework Programme. Switzerland: RWTH and EMPA; 2013.
[81] Leuenberger M, Frischknecht R. Life cycle assessment of battery electric vehicles and concept cars. Report, Uster, Switzerland: ESU-Services Ltd; 2010.
[82] Boyden A. The environmental impacts of recycling portable lithium-ion batteries; [BSc Thesis]. Australian National University; 2014.
[83] GREET 2016. Argonne, IL: Argonne National Laboratory; 9700 S. Cass Avenue, Building 362, Argonne, IL 60439-4844 USA2016.

Further Reading
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
Appl M. Complete ammonia production plants. Ammonia: principles and industrial practice. Wiley-VCH Verlag GmbH; 2007. p. 177–204. http://dx.doi.org/10.1002/
9783527613885.ch05
Appl M. Future perspectives. In: Ammonia: principles and industrial practice. Weinheim: Wiley-VCH Verlag GmbH; 2007. p. 245–9.
Bartels J.R. A feasibility study of implementing an ammonia economy [Graduate theses and Dissertations]. Iowa State University; 2008.
Bicer Y, Dincer I, Vezina G, Raso F. Impact assessment and environmental evaluation of various ammonia production processes. Environ Manage 2017;59(5):842–55.
doi:10.1007/s00267-017-0831-6.
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
Bicer Y, Dincer I. Performance assessment of electrochemical ammonia synthesis using photoelectrochemically produced hydrogen. Int J Energy Res. 2017; doi:10.1002/
er.3756.
Christensen CH, Johannessen T, Sørensen RZ, Nørskov JK. Towards an ammonia-mediated hydrogen economy? Catal Today 2006;111:140–4.
Dincer I., Zamfirescu C. Methods and apparatus for using ammonia as sustainable fuel, refrigerant and nox reduction agent. Patent Nr. US20110011354 A1; 2009.
Duynslaegher C, Contino F, Vandooren J, Jeanmart H. Modeling of ammonia combustion at low pressure. Combust Flame 2012;159:2799–805.
Feng T, Lü L. The characteristics of ammonia storage and the development of model-based control for diesel engine urea-SCR system. J Ind Eng Chem 2015;28:97–109.
Garagounis I, Kyriakou V, Skodra A, Vasileiou E, Stoukides M. Electrochemical synthesis of ammonia in solid electrolyte cells. Front Energy Res 2014;2:1.
Jennings JR. Catalytic ammonia synthesis: fundamentals and practice. New York. US: Springer; 1991.
Kim K, Yoo C-Y, Kim J-N, Yoon HC, Han J-I. Electrochemical synthesis of ammonia from water and nitrogen catalyzed by nano-Fe2O3 and CoFe2O4 suspended in a molten
LiCl-KCl-CsCl electrolyte. Korean J Chem Eng 2016;33:1777–80.
Ouadha A, El-Gotni Y. Integration of an ammonia-water absorption refrigeration system with a marine diesel engine: a thermodynamic study. Procedia Comput Sci
2013;19:754–61.
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. Utilization of hydrogen produced from urea on board to improve performance of vehicles. Int J Hydrogen Energy 2011;36:11425–32.
Ammonia 39

Relevant Websites

http://www.ammoniaenergy.org
Ammonia Energy.
https://ammoniaindustry.com/
Ammonia Industry.
https://www.ohio.edu/engineering/ceer/ammonia-electrolysis.cfm
Center for Electrochemical Engineering Research (CEER) – Ohio University.
https://chemengineering.wikispaces.com/Ammonia+production
ChemEngineering.
https://www.canada.ca/en/health-canada/services/publications/healthy-living/guidelines-canadian-drinking-water-quality-guideline-technical-document-ammonia.html
Government of Canada.
http://www.nh3fuel.com/
Hydrofuel Inc.
https://www.ipni.net/
International Plant Nutrition Institute (IPNI).
http://www.nautilus.org/napsnet/napsnet-policy-forum/monia-as-a-fuel-for-passenger-vehicles-possible-implications-for-greenhouse-gas-reduction-in-korea/
Nautilus Institute.
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.worldfertilizer.com/
Palladian Publications Ltd.
http://www.potashcorp.com/
PotashCorp.
http://www.spg-corp.com/clean-energy-power-generation.html
Space Propulsion Group (SPG).
http://www.linde-engineering.com/en/process_plants/hydrogen_and_synthesis_gas_plants/gas_products/ammonia/index.html
The Linde Group.
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.
http://www.yara.com/
Yara.

Das könnte Ihnen auch gefallen