Download as pdf or txt
Download as pdf or txt
You are on page 1of 84

Teesside University

School of Computing, Engineering and Digital


Technologies

Research Project

2022-2023

AYOB KHALID ABDULLAH AL-LUAHIBI

B1781241@live.tees.ac.uk

Techno economic analysis of hydrogen production


system

Supervisor: Dr. Ahmed, Tariq

1
Abstract

It is anticipated that hydrogen will have a crucial role in shaping the


energy systems of the future. This paper presents a comprehensive
techno-economic analysis and life cycle assessment of various hydrogen
production systems, steam methane reforming, coal gasification,
biomass gasification, and water splitting technologies. The objective is to
identify the most promising and sustainable method for hydrogen
production. The economic feasibility and environmental impact of each
technology have been assessed. The results show fossil fuel-based
hydrogen production has the lowest cost, but it is not sustainable.
Biomass gasification is considered a renewable source and provides a
promising production cost. Electrolysis has high costs, especially if it
uses renewable energy sources, which results in the theft cost. Nuclear
based hydrogen provides competitive production costs and is expected
to be the driver for hydrogen production in the future. There are various
aspects of the carbon footprint for different technologies. The carbon
footprints from steam methane reforming and coal gasification are 11.98
kg CO2/kg H2 and 23 kg CO2/kg H2, respectively. Employing CCS can
capture 90% of carbon emissions. Biomass gasification provides a
promising way that depends on renewable biomass feedstock, and the
emission of carbon with CCS is 3.54 kg CO2/kg H2. Renewable
electrolysis has the highest cost with low emissions (1-2 kg CO2/kg H2),
but it is sustainable with very low carbon emissions. Nuclear-based
hydrogen has very low carbon emissions (0.45-2 kg CO2/kg H2).

2
1 Table of Contents
1Table of Contents ................................................................................................. 3

2Introduction........................................................................................................... 6

2.1Overview: ....................................................................................................... 6

2.2Hydrogen economy: .................................................................................... 10

2.2.1Hydrogen production:............................................................................ 10

2.2.2Hydrogen storage:................................................................................. 11

2.2.3Hydrogen distribution: ........................................................................... 13

2.2.4Hydrogen application: ........................................................................... 14

2.3Aim and objectives ...................................................................................... 17

2.4Scope: .......................................................................................................... 17

2.5Contribution to knowledge: .......................................................................... 17

3Literature Review ............................................................................................... 18

3.1Hydrogen production technologies: ............................................................ 18

3.1.1Natural Gas Steam Methane Reforming (SMR): ................................. 19

3.1.2Coal Gasification: .................................................................................. 22

3.1.3Biomass gasification: ............................................................................ 24

3.1.4Water splitting: ....................................................................................... 27

3.1.5Methane pyrolysis (Turquoise H2):....................................................... 31

3.2Environmental effects of CO2 sequestration: ............................................. 32

4Methodology ....................................................................................................... 33

4.1Literature review: ......................................................................................... 33

4.2Economic Analysis:...................................................................................... 33

4.3Life Cycle Assessment (LCA):..................................................................... 33

4.4Carbon Capture Costs Analysis: ................................................................. 33

4.5Data Interpretation and Conclusion: ........................................................... 34

5Economic Evaluation ......................................................................................... 35

3
5.1Cost parameters involved:........................................................................... 35

5.2Hydrogen production cost: .......................................................................... 35

5.2.1Steam methane reforming: ................................................................... 35

5.2.2Coal gasification: ................................................................................... 38

5.2.3Biomass gasification: ............................................................................ 40

5.2.4Water splitting: ....................................................................................... 42

5.3Uncertainty: .................................................................................................. 45

6H2A Model .......................................................................................................... 48

6.1Natural gas with CO2 sequestration model: ............................................... 48

6.2Coal gasification with CCS model: .............................................................. 49

6.3Biomass gasification model no CCS: .......................................................... 50

6.4Grid electrolysis: .......................................................................................... 51

7Life Cycle Assessment ....................................................................................... 53

7.1Steam Methane Reforming: ........................................................................ 53

7.2Coal Gasification:......................................................................................... 55

7.3Biomass gasification: ................................................................................... 56

7.4Water splitting: ............................................................................................. 58

7.4.1Wind electrolysis: .................................................................................. 58

7.4.2Electrolysis integrated with PV power: ................................................. 60

7.4.3High temperature electrolysis for hydrogen production via nuclear


energy:61

7.5Environment impact assessment: ............................................................... 62

7.6Carbon capture storage:.............................................................................. 65

7.6.1Steam methane reforming: ................................................................... 65

7.6.2Coal gasification .................................................................................... 67

7.6.3Biomass gasification: ............................................................................ 68

7.6.4Electrolysis: ........................................................................................... 68

4
8Result and Discussion ....................................................................................... 69

8.1Production Cost: .......................................................................................... 69

8.2Sensitivity: .................................................................................................... 71

8.2.1Steam methane reforming: ................................................................... 71

8.2.2Coal gasification: ................................................................................... 72

8.2.3Biomass gasification: ............................................................................ 73

8.2.4Electrolysis: ........................................................................................... 73

8.2.5Nuclear hydrogen production:............................................................... 74

8.3Environmental Impact: ................................................................................. 74

9Conclusion And Recommendation .................................................................... 77

5
1 Introduction
1.1 Overview:
With global energy entering a new period of energy concern and multiple
options for energy sources, there is worldwide agreement to regulate
greenhouse gas emissions, posing the issue of determining strategies to reduce
emissions and supporting sustainable energy sources. The expected growth in
global energy consumption, as well as the economic and geopolitical
ramifications of potential oil supply shortages, have been important factors in
the debate over future energy supply. That’s why hydrogen has received
attention as the future clean energy fuel. According to the International Energy
Agency (IEA), the upcoming decade will see an increase of 50% in global
energy demand, with fossil fuels continuing to dominate the energy sector.
There are two important factors for the future energy industry: supply security
and climate change. This raises the problem of determining the best strategy to
reduce emissions while delivering energy needs, which has an effect on global
policymaking. The transport sector uses 18% of energy and contributes 17% of
global CO2 emissions; until 2030, this sector will be responsible for 20% of
energy consumption and green gas emissions. Since oil continues to be the
primary source of energy for the world and accounts for more than 95% of
transportation energy demand, any problems with oil supply distribution will
result in higher prices for the industry, which is almost entirely dependent on oil
(Ball and Wietschel, 2009). Data reveal global hydrogen consumption was 120
Mt/a in 2020 and is suggested to increase to 530 Mt/a by 2050, with over 70
Mt/a of hydrogen used in oil refineries and 39 Mt/a used in ammonia production.
Additionally, around 50 million metric tonnes per year (Mt/a) of hydrogen are
employed in carbon-rich gas mixtures. Methanol synthesis processes consume
14 Mt/a of hydrogen. while the steel production process consumes 5 Mt/a.
Notably, the energy sector consumes 30 Mt/a for power generation and heat
(Hermesmann and Müller, 2022).

There has been lots of study on using hydrogen as a fuel; it can be employed in
the automotive sector with fuel cells. It has an advantage due to the lack of
carbon in hydrogen fuel, which reduces gas emissions, so there will be no CO2
emissions at the end of combustion. A study estimated that hydrogen-powered
vehicles could significantly impact the reduction of CO2 emissions, predicting a
6
decrease of 35% by 2040. Furthermore, if this clean energy source is utilised in
public transportation, the reduction in CO2 emissions could be as high as 40%.
On average, this would translate into a substantial decrease of 44.8 grammes of
CO2 emissions for every kilometre travelled (Efstathios E., 2014). By 2050,
hydrogen will contribute approximately 10% of global energy usage, derived
from low-carbon sources, as delineated by the International Energy Agency for
the global energy sector.

The number of industries investigating the use of low-carbon hydrogen has


increased, along with the number of nations setting high objectives for reducing
greenhouse gas emissions. The 195 signatories to the 2015 Paris Agreement
committed to increasing efforts to reduce emissions towards a net zero target
across all sectors by the end of the century. In 2018, the Intergovernmental
Panel on Climate Change estimated that CO2 needed to maintain a global
temperature rise of 1.5 °C to reach zero emissions by 2050 (The Future of
Hydrogen, 2019).

The falling cost of renewable energy resources is driving hydrogen potential


upward, and as the costs of solar and wind fall, their contribution to the future
primary energy mix is expected to grow. The fluctuation of solar and wind power
generation is a problem of large proportions. Several countries have ambitions
to use energy coming from low-carbon sources to achieve the target of 100%:
South Australia has set a goal to achieve it by 2025; Fukushima Prefecture by
2040; Sweden by 2040; California aims to realise this objective by 2045; and
Denmark's target year is 2050. The European Union's has established an
ambitious goal to reduce emissions by 80–95% by 2050 relative to the metrics
from 1990. This commitment requires a shift away from carbon in electricity
generation and integrating renewable energy resources (The Future of
Hydrogen, 2019).

Hydrogen is now widely regarded as a prospective and ecologically sustainable


alternative to conventional fuels, which include gasoline, diesel, heating oil, and
natural gas, spanning both transport and non-transport sectors. Hydrogen has a
remarkable degree of versatility as a secondary energy carrier, owing to its
ability to be derived from a diverse range of basic energy sources that are

7
abundantly accessible. These primary energy sources encompass coal, natural
gas, biomass, and other forms of waste materials.

Hydrogen is an extremely clean energy carrier; its combustion produces only


water, and it also has a high energy density by mass, which makes it attractive
for decarbonization, especially in the transport and industrial sectors. It is used
in fuel cells and combusted in turbines to generate electricity (netl.doe.gov).
Utilising hydrogen as a fuel source comes with significant environmental
benefits as it doesn't emit greenhouse gases, particulates, sulphur oxides, or
contribute to ground-level ozone. Moreover, when harnessed in a fuel cell, the
sole product of the process is water, further underscoring its environmental
advantages (The Future of Hydrogen, 2019).

In 2019, 70 Mt of hydrogen were produced, with natural gas accounting for 76%
and coal contributing 23%. This production level consumes 205 billion m3 of
natural gas and 107 Mt of coal. The cumulative CO2 emissions from the United
Kingdom and Indonesia are estimated at 830 Mt CO2/year from hydrogen
production. Currently, electrolysis accounts for about 2% of worldwide hydrogen
production; there is potential for electrolysis to create additional low-carbon
hydrogen (The Future of Hydrogen, 2019).

Hydrogen production costs have significant regional variations that influence


fossil fuel prices, electricity prices, and carbon prices. Reforming natural gas
without carbon capture is the most economical method. With nuclear power.
Regions with nuclear power plants may find electrolysis an attractive option.
Most hydrogen generated from renewable sources is more costly in most
regions, ranging from 2.72 to 7.25 £/kg H2, as documented by the International
Energy Agency in 2021. Fig. 1 shows The cost of producing hydrogen differs by
region and depends on the energy sources employed. Hydrogen can be derived
from a variety of low-carbon resources, such as renewable electricity, biomass,
and nuclear power. Additionally, hydrogen can be produced from fossil fuels in a
low-carbon manner if the process is paired with carbon capture technology.

8
Fig. 1. Cost of Hydrogen at different region (Yukesh Kannah et al., 2021).

The current challenge is to meet growing energy demand without polluting the
environment. As shown in Figure 2, the annual demand for H2 is rising,
countries are transforming towards eco-friendly energy to mitigate the
environmental impact associated with energy consumption (Yukesh Kannah et
al., 2021).

Fig. 2. Global hydrogen demand.

9
It is important to find ways to increase H2 production to meet future energy
demand.

1.2 Hydrogen economy:


Hydrogen is gaining popularity as an energy carrier because it has an effective
energy exchange, can be produced from a variety of sources, has a higher
thermal value than conventional fuel, and produces only water as a byproduct.

It is seen as an energy carrier for the future to mitigate carbon emissions; it can
produce heat and power for houses and industries; and it can be used as an
alternative fuel for cars to reduce environmental impact. Hydrogen production
from renewable sources provides a stable energy supply during the year. The
current hydrogen production from steam methane reforming is not sustainable
because of the depletion of natural gas and CO2 production during the
reforming process. Renewable resources must be utilised for sustainable
hydrogen production. Hydrogen is produced from renewable and non-
renewable sources; non-renewable sources dominate. Producing from natural
gas is likely to dominate because it offers low-cost production. The
commercialization of solar methods is expected in the distant future due to the
high expenses associated with photovoltaics and equipment. Coal gasification
is favourable in the US due to the high availability of coal (Mah et al., 2019).

1.2.1 Hydrogen production:


hydrogen is produced from renewable and non- renewable source, the non-
renewable fossil fuels are dominating on hydrogen production 96%, and 4%
produced from electrolysis of water (Mah et al., 2019). the following most used
technology:

1- Steam methane reforming: natural which is composed of methane is used as


feedstock, and it reacts with high-temperature steam to produce hydrogen and
carbon monoxide, as well as a certain quantity of carbon dioxide.

2- Coal gasification: Coal as feedstock reacts with O2 and steam under high
pressure and temperature to produce CO and H2 as syngas. Impurities are
eliminated before proceeding to the next stage.

10
3- Biomass gasification: it use biomass any carbon based substance as
feedstock, it converted to H2, CO, and CO2 at high temperature with steam or a
limited amount of oxygen.

4- Electrolysis: Cathode and anode submerged in electrolyte are used to


generate H2 and O2 from water using direct current.

Natural gas reforming continues to dominate because it is the most cost-


effective method, but it is not sustainable due to its high carbon emissions and
reliance on fossil fuels as feedstock. Biomass gasification is anticipated to be
commercialised in the middle term; it utilises biomass, which is a renewable
source.

Producing hydrogen from renewable sources like wind turbines and solar
panels is still expensive. It is expected to be commercial in the long term. While
electrolysis uses grid electricity, it depends on electricity prices.

Coal gasification with CCS is favourable in the United States due to coal
availability.

1.2.2 Hydrogen storage:


The transport and storage of hydrogen have difficulties due to its low density
per volume. The strategies for hydrogen storage are divided into four categories
in Table 1.

Table 1. Hydrogen storage methods (Mah et al., 2019).

Type of storage Description Challenges

Compressed gas Typically, hydrogen is stored in -Large volume required for


high-pressure gas cylinders storage.
with optimal operating
pressures of 200 bar.

For extensive storage needs,


hydrogen gas is often
pressurized in subterranean
salt caverns to pressures of
200 bar or potentially higher.

11
Cryogenic liquid At exceedingly low The process of transforming
temperatures, specifically hydrogen into its liquid state is
below -253 degrees Celsius, energy-intensive, accounting
liquid hydrogen can be for 30% of the total energy
conserved. contained in the hydrogen.
The storage tank is thermally
insulated and maintained The materials required to
under vacuum conditions. construct the storage tanks for
hydrogen are costly.
Compared to its compressed
gaseous form, hydrogen can Evaporation losses occur in
be stored at a much higher these storage tanks, varying
density in its liquid state. between 0.1% to 1% per day.
Physical storage in hydrides Certain metals or alloys can Optimal storage capacity is
absorb hydrogen, creating typically attained at extremely
metal hydrides that release the low, cryogenic temperatures.
hydrogen upon heating.
It's essential to identify
Metal-organic frameworks, adsorbents that possess higher
carbon-based and other adsorption enthalpies.
nanostructures, and clathrates
are examples of such
materials.
This method of storage is safe,
quick, and reversible under
moderate conditions.
Chemical storage in hydrides Hydrogen forms a chemical -Irreversibility.
bond with the storage
materials. -Impurities absorption.
Metal hydrides, formic acid,
carbohydrates, ammonia, -Reaction kinetics and cost
synthetic hydrocarbons, and
liquid organic hydrogen
carriers are potential materials
for this process.
Compared to metallic hydrides,
these materials could
potentially achieve higher
energy densities.

12
For large-scale hydrogen storage, underground storage and cryogenic liquid
storage are the most feasible strategies. Hydrogen liquefaction is energy-
consuming and used only when there is downstream utility to benefit from the
availability of liquid hydrogen, such as its use in rocket fuel or maritime
transportation. For automobiles and energy storage, there is no need for
hydrogen liquefaction. The most suitable hydrogen storage for hydrogen-
powered vehicles is compressed gas at 700 bar. Hydrogen is stored at between
875 and 1000 bar and -40 °C in refuelling stations. During dispensing, additional
precooling is required to permit a quick replenishment time (3-5 minutes)
without overheating the storage tank.

Underground storage provides the highest hydrogen storage; salt caverns are
promising for storage because they're unreactive with hydrogen. Natural gas-
depleted reservoir or aquifer formations are less favourable for hydrogen
storage because of the possible interaction between hydrogen and minerals or
microorganisms in these formations.

Chemical or physical hydrogen storage in hydrides represents an optimistic


future option. However, there are difficulties that must be addressed before it
can achieve wide viability. There are three prerequisites for an optimal hydrogen
storage material, none of which have yet been unequivocally met: it should
have a high capacity for hydrogen storage, allow reversible charging and
discharging cycles at temperatures between 70 and 100°C, and facilitate quick
charging and discharging kinetics with the smallest possible energy barriers
(Mah et al., 2019).

1.2.3 Hydrogen distribution:


Various transportation methods are employed based on the volume of the
product. Trucks and trailers are used for transporting liquified hydrogen and
gaseous hydrogen in vessels; for gaseous hydrogen delivery, pipelines are
used. Trailers are used to transport hydrogen in small volumes over a short
distance not exceeding 200km while carrying compressed hydrogen (200 bar).
Liquid hydrogen trailers are most effective for long-distance transportation. The
process of liquefying hydrogen requires energy, and cryogenic liquid insulation
is expensive. Conventional transportation methods have limited capacity and
difficulty managing hydrogen; for future hydrogen distribution systems, pipelines

13
are preferable. Therefore, pipeline distribution is preferable for large-scale
production, whereas vehicle transport is appropriate for small-scale production
(Mah et al., 2019).

1.2.4 Hydrogen application:


The chemical industry has used hydrogen to produce ammonia and fertilisers.
In addition, it is used in the refining industry as a petrochemical in hydrocracking
and desulfurization procedures. The production and fabrication of metals, the
synthesis of methanol, and food processing are additional applications of
hydrogen. According to Mah et al. (2019), the global hydrogen production
market is projected to reach 147.85 billion pounds by 2023.

Hydrogen produces energy through two methods. The first requires controlled
combustion of hydrogen, whether in gaseous or liquified form. This combustion
generates energy as heat and results in water vapour as a byproduct. Second,
hydrogen reacts with oxygen in fuel cells, producing electricity along with water
and heat. Its operation is characterised by silence and efficiency, with stationary
parts. As a pure source, it can reduce carbon emissions in a variety of sectors,
including transportation, heating, and energy storage, as shown in Table 2 (Mah
et al., 2019).

Table 2. Applications of hydrogen as energy carrier.

Application Description Examples


Automobiles Hydrogen can be used directly Modification of spark-ignition
as a fuel in an internal gasoline-powered internal
combustion engine (ICE), or it combustion engine of a
can be used to generate Volkswagen Polo 1.4 to enable
electricity for fuel cell vehicles it to operate on hydrogen. This
(FCVs). modification allows the vehicle
to reach a maximum speed of
140 km/h and have sufficient
power reserve to navigate
standard urban routes and
inclines of up to 10%.

Currently available hydrogen


fuel cell vehicles in the market

14
include models like the Toyota
Mirai, Hyundai Nexo, and
Honda Clarity.
Residential and industrial H2 can be utilised to generate Adding hydrogen to natural
heating thermal energy. gas, up to a concentration of
30 mol%, can reduce CO2
emissions by up to 18% from
domestic appliances such as
burners and boilers, without
any need for modification.

A refinery's fuel gas can


potentially be replaced with
hydrogen by utilizing simulation
models on a fired heater
design.

Micro CHPs (Combined Heat


and Power) units powered by
hydrogen-fueled fuel cells can
be deployed in buildings and
industries to optimise energy
recovery.
Energy storage Hydrogen serves as a valuable Hydrogen fuel cells can be
storage medium for surplus employed as standby power
energy, discharging it as generators for mobile homes
needed to equilibrate supply when the supply from wind and
and demand of electricity. solar energy falls short of
demand.
Benefits of hydrogen fuel cells
include high energy density
and scalability ranging from
kilowatts to multiple
megawatts.

The hydrogen used in chemical, heating, energy, and transport. Table 3


summarises the various ways of hydrogen application.

15
Table 3. Various hydrogen application.

Application Description
Hydrocracking Converts heavy oil products into lighter ones using hydrogen.
Hydrotreating Removes contaminants such as S and N from refinery
feedstocks with the help of hydrogen.
Ammonia Hydrogen is a crucial component in the industrial production of
Production ammonia which has diverse applications like in fertilizers,
chemical synthesis, explosives, fibre and plastics, refrigeration,
pharmaceuticals, pulp and paper, mining and metallurgy,
cleaning.
Hydrogenation Involves the addition of hydrogen to vegetable oils under the
of Fats presence of a catalyst.
Iron Production Hydrogen used as a reductant substitute in the production of
(DRI) Direct Reduced Iron (DRI).
Methanol Hydrogen is used to produce methanol, a key component in the
Production chemical industry and fuels.
Heat Treatment Hydrogen is used in manufacturing processes for aerospace,
Processes automotive, and petrochemical industries.
Plastics Future route to manufacture PET from hydrogen and captured
Production CO2.
Forming Gas Hydrogen is used in catalyst regeneration, Photographic
and hypersensitization, semiconductor applications etc.
Semiconductor
Glass Hydrogen used in glass manufacturing though more research is
Production needed for optimization.
Cooling of High heat capacity of hydrogen makes it suitable for cooling
Generators applications.
Cement Hydrogen has been used to substitute some of the fossil fuels in
Production cement production.
Transportation Hydrogen fuel cells are used for powering vehicles. This
includes buses, trains, and airplanes. Also, hydrogen-derived

16
fuels like ammonia and methanol can be used in the maritime
sector.
Heating Hydrogen could be used to decarbonise heating, replacing fossil
fuels.
Energy Hydrogen is used as an energy vector and can provide flexibility
in the power grid.

1.3 Aim and objectives


The aim of this study is to identify hydrogen production systems to evaluate the
technical and economic viability of different hydrogen production technologies.
The aim will be achieved by following objectives:

• To conduct a comprehensive review of relevant literature, research


papers, and industry reports to gather information on hydrogen
production technologies.
• To collect technical and economic data for different hydrogen systems.
• To compare the collected data and identify the feasibility of each
production method.
• To carry out a life cycle assessment of the different hydrogen production
systems identified above.

1.4 Scope:
This study focuses on the economic feasibility of hydrogen production systems.
It will consider all aspects of hydrogen production, including different
technologies, feedstocks, and processes.

1.5 Contribution to knowledge:


This study aims to contribute to the knowledge of hydrogen production systems
by providing a comprehensive analysis of the most economic hydrogen
production technologies. The study will help to improve understanding of the
cost drivers and the key factors that impact the economic viability of these
technologies.

17
2 Literature Review
This chapter looks at the technical aspects of hydrogen production
technologies.

2.1 Hydrogen production technologies:


Hydrogen can be derived from various methods, such as fossil fuels, biomass,
water, electrolysis, or a combination of them. Producing hydrogen consumes
275 Mtoe of energy (2% of global energy demand), and natural gas is the
primary source for hydrogen production, particularly through steam methane
reforming. It produces about 70 million metric tonnes of hydrogen (about 76% of
the overall hydrogen production), consuming 6% of global natural gas demand.
Coal gasification produces 23% of overall hydrogen production, consuming 107
Mt of coal, which is 2% of global coal demand. Hydrogen production
technologies can be divided into four categories: thermal process,
electrochemical process, biological process, and thermochemical process.
Hydrogen is referred as a colour based on production technology, as shown in
Table 4 (Massarweh et al., 2023).

Table 4. Comparison between the different types of H2 (Massarweh et al.,


2023).

Type Primary source/process CO2 emissions


Brown H2 Fossil fuels (coal) CO2 is released into the
environment
Gray H2 Fossil fuels (natural gas) CO2 is released into the
environment
Blue H2 Fossil fuels (natural gas) CCS techniques are
employed
Green H2 Water electrolysis using Zero or minimal CO2
electricity produced from emissions
renewable energy
Turquoise H2 Fossil fuels (mostly from Solid carbon production
natural gas)
Gold H2 Biologically produced Carbon neutral
from residual

18
hydrocarbons using
microbes in the
subsurface (i.e., depleted
oil reservoirs)
Purple H2 Water electrolysis based Zero or minimal CO2
on nuclear energy emissions

2.1.1 Natural Gas Steam Methane Reforming (SMR):


Steam methane reforming is a dominant technique for producing hydrogen.
SMR is expected to retain its position in large-scale hydrogen production due to
its cost-effectiveness. Integrating carbon capture can reduce up to 90% of the
associated emissions, reducing the environmental impact (The Future of
Hydrogen, 2019).

2.1.1.1 Process performance without CO2 emissions reduction (grey


hydrogen):
Conventional steam methane reforming. The process flow diagram presented in
Fig. 3 (Katebah, Al-Rawashdeh, and Linke, 2022).

Fig. 3. Diagram of the SMR process for the production of grey hydrogen.

The pre-reforming procedure is fed natural gas and water steam, where the
natural gas is transformed into methane. This process prevents soot and
enhances efficiency. Additional steam is added before the mixture enters the
main reactor. This stage is an endothermic reaction, producing carbon
monoxide and hydrogen (eq. 1) (Katebah, Al-Rawashdeh and Linke, 2022).

CH4+ H2O → CO + 3H2 (1)

19
Radiant and convective are the two components of the reactor. The radiant has
tubes in addition to a burner, which is fueled by natural gas and PSA (pressure
swing adsorption) exhaust gas. The flue gas is channelled to the convective
section to recover heat. To further hydrogen production yield, exothermic water-
gas-shift is employed; this reaction transforms carbon monoxide into hydrogen
and carbon dioxide (eq. 2).

CO + H2O → CO2+3H2 (2)

The water-to-gas transformation favours low temperatures, although reaction


rates are quicker at high temperatures. Many industrial plants set up dual-stage
high-temperature and high-temperature shift reactors. After shifting, the syngas
is exposed to separation to produce pure hydrogen. PSA is employed in the
majority of plants (85%) for the separation and purification of hydrogen due to
its low capital and operating costs. The pressure-swing adsorption process
consists of two adsorbers that run under cyclic operation. Hydrogen flows
through adsorption columns; impurities and CO2 get adsorbed into internal
surfaces. Reducing column pressure to atmosphere pressure leads to the
desorption of impurities, resulting in high pure hydrogen production. The PSA
gases, which consist of unreacted methane, hydrogen, and CO2, burn with
additional natural gas as fuel for the reactor furnace, ensuring an efficient use of
process resources (Katebah, Al-Rawashdeh, and Linke, 2022).

2.1.1.2 Process integration of SMR with CO2 capture and compression


(blue hydrogen):
Carbon capture and storage (CCS) is used at an SMR plant to produce blue
hydrogen in order to reduce carbon dioxide emissions. CCS is comprised of a
series of stages, including separating CO2 from energy-related industrial
sources, compressing it, and transporting it for long-term storage. The energy
needed for CO2 capture and compression, as well as transport and storage
expenses, contribute to an increase in the cost of producing hydrogen.

In SMR plants, 75% of the CO2 emissions come from the reforming process
and the water-gas shift reaction. In the reformer furnace, the combustion of PSA

20
combustion gases and additional combustible gas produces the remaining 25%.
The capture strategy can be implemented at three distinct points: from shifted
syngas as it exists in the low-temperature water-gas-shift reactor; from PSA,
steam contains unreacted methane, H2, and CO2; and from the reformer
furnace, which results from burning the fuel and tail gas. These options are
represented in the diagram in Fig. 4 (Katebah, Al-Rawashdeh, and Linke, 2022).

Fig. 4. CO2 removal location options in the SMR plant.

The position of CO2 capture is determined by various factors, including, gas


stream pressure, concentration of CO2, and impurities existence (Massarweh et
al., 2023).

Fig. 5. Diagram of a typical amine carbon capture unit's process flow.

21
Fig. 5 represents the CO2 removal process. This process requires an absorber
and a regenerator. In the absorber, the CO2 is captured by a lean solvent,
producing a CO2-rich solvent. This solvent undergoes heating and
depressurization before being sent to the regenerator to separate CO2. The
refreshed solvent is recycled back to the absorber. The CO2 from the amine
unit is compressed and stored. Sulfinol-X and ADIP-X have been identified as
the leading solvents for post-combustion carbon capture. due to their loading
capacity, chemical resilience, minimal corrosiveness, vapour pressure, and
energy demands. However, monoethanolamide (MEA) remains the
conventional choice for CO2 capture from flue gas (Katebah, Al-Rawashdeh,
and Linke, 2022).

2.1.2 Coal Gasification:


Hydrogen production by coal gasification has been used for several decades.
There are around 130 coal gasification plants around the world, with 80% in
China. Coal gasification produces CO2 emissions of about 19 tCO2/tH2, which
is twice the amount of natural gas reforming, so carbon capture technology is
used to mitigate gas emissions. However, coal gasification brings challenges
due to a low hydrogen-carbon ratio (0.1:1 from coal vs. 4:1 from methane) and
high levels of impurities in the feedstock (sulphur, nitrogen, and minerals) (The
Future of Hydrogen, 2019).

22
Coal gasification is currently a commercial method for producing hydrogen for
the chemical sector. This method has been successfully adopted in areas where
natural gas is not available, Müller-Langer et al. (2007).

Fig. 6. Gasification of coal—conventional system components

Fig. 6 Shows the process. The procedure of coal gasification involves the high-
temperature interaction of pulverised coal with air and steam in a gasifier,
forming syngas consisting of H2, CO, CO2, traces of CH4, and residual steam.
The raw gas contains contaminants like carbonyl sulphide and hydrogen
sulphide. The raw gas is treated to generate two streams, one enriched in
hydrogen and the other in CO2.

Step 1: The gasification process transforms coal into syngas, heated up to 1800
°C. The syngas is comprised of CO, H2, CO2, and minor quantities of gases
and particulates.

Step 2: Cleaning: The syngas is then cooled and purified to eliminate any other
gases and particulates. This leaves only CO,H2, and CO2. During the syngas
purification stage, mercury, trace contaminants, and particulate matter are
removed.

23
Step 3: Shifting : After the cleaning stage, the syngas is processed in a shifting
reactor, where the carbon monoxide reacts with steam to form CO2 and
additional hydrogen.

Step 4: Purification: Post-shifting, the syngas is divided into two streams:


hydrogen and carbon dioxide. The purified hydrogen is prepared for use, and
the carbon dioxide is captured and designated for sequestration.

Step 5: Usage: The pure hydrogen stream is ready for use in a variety of
applications; it can be combusted in a gas turbine to produce electricity or via
fuel cells. It can be employed in an internal combustion engine or used in
chemical production.

With carbon capture and storage, carbon dioxide emissions can be mitigated to
near zero. For example, a coal plant with 500 MW produces 444 tonnes of CO2
per hour; deploying carbon capture mitigates emissions to 36 tonnes, which is
reduced by 90%. Carbon capture can be applied to electrical power plants.

Coal gasification plants are more complex than steam SMR plants due to:

(i) The process requires extensive feedstock preparation, including coal


grinding, drying, and creating coal slurry, depending on the gasifier type.

(ii) The need for an air separation unit that is used to produce oxygen for an
oxygen-blown gasifier.

(iii) The need for multi-stage cleaning of raw syngas, which includes the
removal of particulate matter, dust, and sulphur.
(iv) The requirement for further syngas processing beyond the conversion of
CO.

Small-scale coal plants are not feasible due to their complexity, which requires
high capital costs (Muellerlanger et al., 2007).

2.1.3 Biomass gasification:


Biomass gasification is the process that converts organic materials, wood,
agricultural residues, and other biomass into syngas. This is achieved by
heating biomass material to a high temperature (726 °C). This material

24
undergoes partial oxidation, resulting in the gas charcoal. This charcoal is
decomposed to produce hydrogen, carbon monoxide, carbon dioxide, and
methane (Ni et al., 2006).

Biomass + heat + steam→H2 + CO + CO2 + CH4 + light and heavy (3)


hydrocarbons + char

The produced gas can be reformed by steam to produce hydrogen and can be
improved by using the water-gas shift reaction.

A significant issue in biomass gasification is managing the tar formation that is


generated during the process. It is an undesirable product that causes issues.
To minimise tar production, there are three strategies: designing gasifiers
appropriately, implementing effective control and operation measures, and using
additives and catalysts. Biomass gasification faces another challenge of ash
formation leading to deposition, sintering, slagging, fouling, and agglomeration
within the reactor. Strategies such as fractionation and leaching are used to
curtail ash production. Fractionation is an effective method to remove ash and
might deteriorate the quality of the remaining ash. Alternatively, leaching can
remove the inorganic portion of biomass and enhance the quality of the
remaining ash (Ni et al., 2006).

Fluidized bed (FB), fixed bed, and entrained flow (EF) gasifiers are among the
types that have undergone thorough research and practical testing for
commercial applications (Salkuyeh, Saville, and MacLean, 2018).

FB gasifiers have the advantage of reducing the time frame for temperature
distribution. They are flexible with feedstock variation and composition, which
makes them a viable choice for large-scale operations. Another option, an
indirect circulating FB gasification system, has been innovated to prevent
syngas combustion within the gasification reactor, to prevent dilution of syngas
with nitrogen, and to eliminate the use of high-purity oxygen. FB gasifiers
operate at low temperatures, resulting in low conversion of biomass to syngas.
This requires an additional downstream unit dedicated to reforming tar and
other hydrocarbons to hydrogen and carbon monoxide (Salkuyeh, Saville, and
MacLean, 2018).

25
Entrained flow (EF) gasifiers have a lengthy history of use in coal and petcoke
gasification and are a well-established commercial technology. When using this
technology to gasify biomass, the biomass is partially combusted within the
gasifier, resulting in a higher temperature than FB gasification technology. This
elevated temperature guarantees that the internal tar cracks, resulting in the
formation of syngas without tar. In addition, EF gasifiers can achieve a high
carbon conversion rate in a short amount of time. This reduces the size of the
gasifier, resulting in reduced costs. For the partial combustion of feedstock, EF
gasifiers require high-purity oxygen (Salkuyeh, Saville, and MacLean, 2018).

Figures 7 and 8 provide simplified block diagrams for the FB and EF gasification
processes, respectively, detailing configurations with and without carbon
capture and liquefaction integration. Option (a) shows a configuration in which
the produced CO2 is discharged into the atmosphere. Option (b) shows a
carbon capture system in which all CO2 coming from the syngas (via the Amine
unit) and the CO2 from the exhaust gas stream of the power and steam
generation module are isolated and then liquefied (Salkuyeh, Saville, and
MacLean, 2018).

Fig. 7. Fluidized bed for hydrogen production.

gasification system. Option (a): Without carbon capture, Option (b): With carbon
capture.

26
Fig. 8. Entrained bed for hydrogen production

gasification system. Option (a): without carbon capture, Option (b): with carbon
capture.

2.1.4 Water splitting:


Water is the most abundant element on earth. It is used for hydrogen production
through water splitting, such as electrolysis, thermolysis, and photo-electrolysis.
When renewable energy is used, it paves the way for the cleanest energy
carrier.

2.1.4.1 Electrolysis:
Humanity first produced hydrogen through water electrolysis. The resurgence of
electrolysis was prompted by the energy crisis that occurred throughout the
1970s. Currently, electrolysis is being used on a small scale. Water electrolysis
is used, especially in renewable energy sources, to balance the demand and
availability of solar and wind power (Ji and Wang, 2021).

The process of electrolysis essentially uses direct current; the water


decomposes into hydrogen and oxygen, and the cathode and anode facilitate
the production of oxygen and hydrogen, which can be readily separated into
their respective forms. This process takes place when the cathode and anode
are submerged in water. (Ji and Wang, 2021)

27
2H2O → 2H2 + O2 (4)

AEL and PEM electrolysis are commonly used for commercial purposes and
performed in several tens of degrees; they are the most mature electrolysis
technologies. In alkaline electrolysis (AEL), two electrodes are immersed in a
concentrated alkaline solution composed of KOH. Hydroxide ions, carrying
negative charge, travel from cathode to anode through the porous diaphragm,
allowing only negative ions to pass. This movement results in the production of
water from the hydrogen ions generated at the anode (Ji and Wang, 2021).

Cathode: 4H2O + 4e− → 2H2 + 4OH− (5)

Anode: 2H2O → O2 + 4H+ + 4e− (6)

Fig. 9. Schematic of the operating principle of an alkaline and PEM water


electrolysis cell (Carmo et al., 2013).

This method for producing hydrogen is mostly employed when purity of


hydrogen is necessary. The electrolysis process requires electricity and water; it
requires 9 litres of water to produce 1 kg of hydrogen and a byproduct of 8kg of
oxygen. It takes twice as much water as is used in SMR; hydrogen production is
around 7 MtH2 from electrolysis, which is 1.3% of the global water consumption

28
(345 million m3 of water for 52 MtH2 from SMR) (The Future of Hydrogen,
2019).

Alkaline electrolysis, proton exchange membrane (PEM) electrolysis, and solid


oxide electrolysis cells (SOECs) are the three primary electrolysis technologies.
Alkaline electrolysis has reached a point of commercial viability. Since the
1920s, it has been used primarily in the fertiliser and chlorine industries. PEM
electrolyser systems were introduced in the 1960s to overcome the drawbacks
of alkaline electrolysis. They use pure water as an electrolyte solution to
eliminate recycling potassium hydroxide in solution. SOECs remain the least
advanced. (The Future of Hydrogen, 2019).

High-temperature electrolysis (HTSE) uses both heat and electricity to produce


hydrogen integrated with a nuclear reactor and to provide an endothermic
reaction to split water into oxygen and hydrogen. Compared to standard
electrolysis HTSE is more efficient because it takes advantage of using high
temperatures to reduce the required electricity.

2.1.4.2 Thermolysis:
Thermolysis, or thermochemical water splitting, involves heating to decompose
water into hydrogen and oxygen. The decomposition process requires a very
high temperature, above 2500 °C. Due to the high energy required that cannot
be provided by a sustainable heat source, several thermochemical water-
splitting cycles have been suggested to reduce the required temperature and
enhance heat. Thermochemical cycles are sequences of chemical reactions to
enhance efficiency and reduce the required temperature. The thermochemical
cycle, which consists of a sequence of chemical reactions occurring at different
temperatures, is promised as a method to convert heat to chemical energy in
hydrogen form (Nikolaidis and Poullikkas, 2017).

-Single-stage water decomposition :

2H2O→2H2+O2 T>2500 °C (7)

-Multi-stage Cu-Cl cycle :

2CuCl2(s)+H2O(g)→CuO*CuCl2(s)+2HCl(g) T=400 °C (8)

29
CuO*CuCl2(s)→2CuCl(l) +0.5O2 T=500 °C (9)

4CuCl(s)+H2O→2CuCl2(aq)+2Cu(s) T=25–80 °C (10)

CuCl2(aq)→CuCl2(g) T=100 °C (11)

2Cu(s)+2HCl(g)→2CuCl(l) H2(g) T=430–475 °C (12)

Novel two-step SnO2/SnO cycle :

SnO2(s)→SnO(g)+0.5O2 T=1600 °C (13)

SnO(s)+H2O(g)→SnO2(s)+H2(g) T=550 °C (14)

Several thermochemical cycles are being investigated for their technical


viability. The Sulphur-Iodine (S-I), Hybrid Sulphur (HyS), Copper-Chlorine (Cu-
Cl), and Magnesium-Chlorine (Mg-Cl) cycles are examples. Electro-
thermochemical cycles have the most potential. In particular, the hybrid sulphur
(HyS) and Copper-Chlorine (Cu-Cl) cycles are the most advanced. El-Emam
and zcan (2019) report that the Calcium-Bromine (Ca-Br) and Magnesium-
Chlorine (Mg-Cl) cycles show promise as medium-temperature cycles.
Thermochemical cycles are integrated with nuclear reactors to produce
hydrogen at large scales.

There are several reactor types: Advanced Pressurised Water Reactor (APWR),
High-Temperature Gas Cooled Reactor (HTGR), Gas Turbine High Temperature
Reactor (GT-HTR), and HTR-PM (High-Temperature Reactor—Pebble Module).

30
2.1.5 Methane pyrolysis (Turquoise H2):
Methane pyrolysis is a method to produce hydrogen in the absence of oxygen
at low emissions. It is an endothermic reaction that requires high temperatures
(800–1600 °C) to break bonds between carbon and hydrogen. It produces
hydrogen and solid carbon. Solid catalysts, molten metals, and molten salts are
among the approaches used to overcome operational challenges and increase
thermal efficiency (Shashank Reddy Patlolla et al., 2023). Pyrolysis is the
process of breaking down molecules of methane using heat. The hydrogen
produced from methane pyrolysis is known as turquoise hydrogen. The process
produced solid carbon as a byproduct, which is easy to store and use in various
industries. Methane pyrolysis reaction (18)

CH4 → C + 2H2 -74.9 kJ/mol (18)

Methane pyrolysis provides a significant reduction in carbon emissions, with a


decrease of about 75% compared to SMR, emitting only 2-3 kg CO2/kg H2
compared to 9–12 kg CO2/kg H2 of SMR. Additionally, a byproduct of solid
carbon is a valuable industrial product that adds additional revenue. However,
process efficiency is contingent on several factors, such as pressure and
catalysts. So, maintaining the reactor at optimal pressure is important for peak
performance. Further extensive research is required, particularly in their techno-
economic and environmental impacts. This technology faces challenges that
need to be addressed before commercialization (Shashank Reddy Patlolla et
al., 2023).

Pyrolysis technology commercialization faces challenges, particularly


concerning safety, dependability of commercial systems, and proper storage of
solid carbon byproducts. While the technology is low-emission, depending on
the heat source, The engagement of academia, industry, and policymakers is
important to realise sustainable hydrogen application across all sectors
(Shashank Reddy Patlolla et al., 2023).

31
2.2 Environmental effects of CO2 sequestration:
Fossil fuel usage main issue is producing carbon dioxide emissions, causing
climate change and global warming. In Qatar, a staggering 96% of all
greenhouse gas emissions is produced from energy industries, including crude
oil and natural gas. As per the Intergovernmental Panel on Climate Change
(IPCC), around 40% of all anthropogenic carbon dioxide emissions are
produced by the top 8,000 emission points. Nations worldwide are deploying
various mitigation strategies to reduce CO2 emissions by 50–90%, to achieve
greenhouse gases stabilization to 450–550 ppm CO2 equivalent in the
atmosphere by mid-century, in alignment with the Paris Climate Accord.
Strategy is require employing carbon capture and storage (CCS) to sequester
greenhouse gas emissions from power generation plants, followed by the
compression and liquefaction of CO2, and its subsequent underground
injection. Carbon dioxide sequestration is widely known as one of the vital
technologies to reduce carbon dioxide and combat global warming. Proper CCS
installation has the potential to lower CO2 concentrations to 450 ppm by the
turn of the century (Massarweh et al., 2023).

Although benefits of CCS, certain concerns are associated with its use.
Foremost among these is the potential risk of accidental CO2 leaks from
pipelines, which can lead to groundwater acidification and underground water
pollution and alter the PH of underground CO2. It has been observed that a
high rate of CO2 injection can stress on geological formations, resulting
fractures. For effective carbon dioxide sequestration, specific criteria need to be
met by the geological formations: they must have a large capacity to store large
CO2 volumes, they must contain seal rock to prevent CO2 from reaching the
surface, they should allow high CO2 injection rates through sufficient
permeability, and they must be deep enough to reserve the stored CO2
securely. Moreover, they must be devoid of leakage channels such as fractures
and faults. Under typical gas reservoir conditions (i.e., temperatures above
31 °C and pressures above 73.8 bar), CO2 acts as a supercritical fluid,
suggesting that the injection of CO2 into deep, depleted gas fields could present
an effective approach for geologic CO2 storage (Massarweh et al., 2023).

32
3 Methodology
To determine the technical and economic feasibility of different hydrogen
production system, the study provides literature review, economic analysis, life
cycle assessment, and carbon capture cost.

3.1 Literature review:


A systematic and comprehensive review of literature, research papers, and
industry reports is conducted to gather data on various hydrogen production
technologies. This review focuses on understanding the economic aspects,
specifically the costs, of these technologies. The search is carried out using
Google Scholar, using keywords such as "hydrogen production technologies",
"economic analysis", "hydrogen economics,", "cost of production", "life cycle
assessment"," carbon capture cost", etc. The most published papers use USD
currency; the exchange rate used in this study is 1 GBP = 1.24 USD (March
2023). The value of money is changed from the year of the published paper to
2023 using the U.S. Inflation Calculator.

3.2 Economic Analysis:


The collected economic data are compared to identify key drivers for the cost of
production across different technologies. Each technology is assessed on
parameters such as initial investment, operational cost, production cost, etc.
The H2A model is utilised for this step. This model is developed by the U.S.
Department of Energy's Hydrogen and Fuel Cells Program, and it provides an
economic analysis for different hydrogen production technologies and an
environmental impact assessment.

3.3 Life Cycle Assessment (LCA):


The LCA data for the hydrogen production technologies are collected from the
reviewed literature. This data highlights the resource consumption and carbon
footprint associated with each technology. It provides insight into the
environmental impacts of the technologies.

3.4 Carbon Capture Costs Analysis:


The costs associated with carbon capture are also collected and analysed. This
is especially relevant for hydrogen production technologies that result in
significant carbon emissions. Including this aspect will ensure a more holistic
understanding of the cost and feasibility of these technologies.

33
3.5 Data Interpretation and Conclusion:
All collected and analysed data are interpreted to draw conclusions regarding
the technical and economic feasibility of each hydrogen production technology.
This is based on their cost-effectiveness, environmental impact, and the key
cost drivers identified.

Table 5. List of used resources for this paper.

Reference Technology Indicator


Collodi et al. (2017) Steam methane reforming Economic
Varbanov et al. (2020) Steam methane reforming Economic
Li et al. (2022) Coal gasification Economic
Salkuyeh, Saville and Biomass gasification Economic
MacLean, (2018)
Kuckshinrichs, Ketelaer, Grid electrolysis Economic
and Koj (2017)
Orhan, Dincer and Rosen, Nuclear based water Economic
(2010) splitting
H2A models SMR/CG/BG/ Grid Economic/LCA
(www.nrel.gov) electrolysis
Spath and Mann, (2000) Steam methane reforming Life cycle assessment
Cetinkaya, Dincer and Coal gasification Life cycle assessment
Naterer, (2012)
Li et al., (2020) Biomass gasification Life cycle assessment
Bhandari, Trudewind and Wind electrolysis Life cycle assessment
Zapp, (2014)
Cetinkaya, Dincer and PV electrolysis Life cycle assessment
Naterer, (2012)
Utgikar and Thiesen, HTE-Nuclear Life cycle assessment
(2006)

34
4 Economic Evaluation
This chapter explores the economic aspects of the production of hydrogen by
various technologies.

4.1 Cost parameters involved:


Various factors that contribute to the design of the plant are considered when
analysing the hydrogen production process's costs. These expenditures consist
of both capital and operating expenses. Capital expenses are one-time
investments, including reactor and land expenditures. Operating expenses
include electricity, labour, water, crop costs, feedstock costs, etc.

Direct capital costs include instruments, piping, controls, and electrical


components, whereas indirect capital costs consist of project construction costs,
supervision fees, contractor fees, and contingencies.

In the steam methane reforming method for hydrogen production, capital costs
primarily include reactor and land costs. Operating costs are comprised of
electricity, water, labour, and raw material costs.

The capital cost of the electrolyser and the electricity cost during operation have
a significant impact on the cost of producing hydrogen via electrolysis. In a
water electrolysis system, the cost parameters include direct capital costs (such
as the electrolyser, storage system, and compressor), fixed operational costs,
material costs, labour costs, variable operating costs, and the costs of steam,
nitrogen, potassium hydroxide (KOH), and deionized water.

For solar and wind turbine electrolyser systems, the capital cost will be higher
due to the expensive installation of solar panels and wind turbines.

4.2 Hydrogen production cost:


Cost is the primary factor influencing the commercialization of technology.

4.2.1 Steam methane reforming:


Steam methane reforming plant produces 215,700 H2 kg/day; the produced
hydrogen cost is 1.42 £/kg H2 (efficiency 76%) without CCS, and LCOH of 2
£/kg H2 (efficiency 70%) with CCS with a 90% capture rate; the CO2 avoidance
cost is 70 £/tonne. The capital cost of a plant without CCS is 192.25 million
pounds; installing CCS increases capital by 78.6% to 343.3 million pounds. The
plant's operating expenses (OPEX) encompass costs related to labour,

35
operations and maintenance (O&M), feedstock, fuel, catalysts, and chemicals.
These costs saw an escalation from £88.87 million annually to £101.19 million.
The most significant contributors to the increased operating cost of the plant
integrated with CO2 capture include feedstock and fuel expenses, maintenance
costs, and the potential revenue losses from electricity sales that are being
used to capture and sequestration processes. Natural gas to hydrogen: 3.3
kg/kg H2, and natural gas price: 9 £/GJ (Collodi et al., 2017).

Varbanov et al. (2020) analysed a steam methane reforming (SMR) plant


located in Northern Norway, capable of producing 450 tonnes of hydrogen per
day. The researchers reported hydrogen production costs of £1.4/kg H2 without
carbon capture and £1.7/kg H2 with carbon emissions capture and storage.
This increase in hydrogen production cost resulted in a CO2 avoidance cost of
£70 per tonne of CO2 avoided. The primary contributor to the CO2 avoidance
cost was CO2 capture and conditioning, accounting for 57% of the total, while
pipeline transport and storage contributed 17% and 26%, respectively. In their
analysis, the price of natural gas was assumed to be 6.25 £/GJ. The results
show feedstock contributes 85% of the total cost and 57% without and with
CCS, respectively.

(Bartels, Pate, and Olson, 2010) they have conducted several studies, and they
found SMR with CCS of capital cost of 286.84 M£ (without CCS 228.8 M£), the
plant capacity of 341,448 kg/day H2, natural gas price of 11.45 £/GJ, the
production cost is 2.87 £/kg H2 (without CCS 2.63 £/kg H2).

Small-scale steam methane reforming produces 500 H2 kg/day; the production


cost is 1.64 £/kg H2 with a CCS capture rate of 90%; the capital cost is 819.225
million pounds; the operating cost is 164,354 £/year, natural gas price is 5.5
£/GJ (Lee et al., 2021).

Table 6. Summarize SMR plants with their production cost and capacities.

Plant capacity 215,700 kg 450,000 kg 500 kg 341,448 kg


H2/day H2/day H2/day H2/day
Capital cost 192.25 Million 228.8 M£
without CCS £

36
Capital cost with 343.3 Million 819.225 286.84 M£
CCS £ Million £
Feedstock cost 9 £/GJ 6.25 £/GJ 5.5 £/GJ 11.45 £/GJ
(Natural gas)
Production cost 1.42 £/kg H2 £1.4/kg 2.63 £/kg H2
without CCS
Production cost 2 £/kg £1.7/kg 1.64 £/kg 2.87 £/kg H2
with CCS H2
Reference (Collodi et al., Varbanov et (Lee et al., (Bartels, Pate,
2017) al. (2020) 2021). and Olson,
2010)

The primary cost determinant is the expense of feedstock. The second is the
cost associated with CO2 capture. The hydrogen production cost is mainly
influenced by natural gas price. This is considerably higher than the capital
expenditure. For larger plants, the feedstock cost constitutes roughly 52% to
68% of the total, whereas for smaller plants, it accounts for about 40%. An
equation developed by Grey and Tomlinson to demonstrate that the cost of
hydrogen is highly sensitive to the price of natural gas as follows:

Hydrogen Cost ($/MMBtu) = 1.27 x NG price ( $/MMBtu) + 0.985 (19)

The equation applicable for plants with a capacity of 236,239 kg/day and a
capital cost of 0.65–0.80 $/SCFD. Penner provides a similar equation for the
cost of hydrogen as follows:

Hydrogen Cost ($/kg) = 0.286 x NG price ($/MMBtu) + 0.15 (20)

This equation did not include plant size and operating assumptions (Bartels,
Pate, and Olson, 2010).

Several factors significantly influence the cost of hydrogen production via SMR,
including the feedstock price, capital cost, operational cost, and whether or not
CCS technology is integrated. Among these, the feedstock price and the CO2
capture cost emerge as the two most critical cost drivers. The implications of

37
these findings suggest a need for strategies to manage feedstock prices and
improve CCS technologies' cost-effectiveness.

4.2.2 Coal gasification:


(Li et al., 2022) have studied coal gasification plants in different provinces in
China and found that hydrogen production costs from coal display significant
regional variations within China due to differing energy prices, with a production
rate of 246,260 kg H2/day. Hydrogen production cost under average electricity
and coal costs is 1.06 £/kg H2 without CCS and 1.58 £/kg H2 with CCS. Capital
cost: 194.3 million pounds; employing CCS increases capital cost to 247 million
pounds. CO2 emissions were reduced from 17.77 kg CO2/kg H2 to 8.75 kg
CO2/kg H2.

Table 7. Coal and electricity prices in China (Li et al., 2022).

Min Ave Max


Coal prices 1.06 £/GJ 2.14 £/GJ 2.87 £/GJ
Electricity prices 44.8 £/MWh 67.5 £/MWh 82.8 £/MWh

Due to differences in energy prices, the regional distribution of production costs


varies. Typically, China rises from the northwest to the southeast. China's
Xinjiang, endowed with abundant coal resources that are easily accessible and
inexpensive, has the lowest hydrogen production costs in China, at £0.75 per kg
of H2 without CCS technology. When CCS technology is implemented, the cost
rises to £1.21 per kg of H2. Guangxi has a relatively high coal price, leading to
the highest hydrogen production costs in the country. Without CCS, the cost is
£1.23 per kg of H2, which is 62% higher than in Xinjiang. Upon integrating CCS
technology, the cost increased to £1.77 per kg H2, marking the highest in
China. Carbon Capture and Storage (CCS) has the potential to substantially
lower carbon emissions; however, its integration significantly increases
hydrogen production costs. When CCS is employed with a coal gasification
plant in China, the cost sees a surge of approximately 49.75%. The avoidance
cost of carbon varied in China's provinces, which was greatly affected by
electricity prices. From west to east China, the regional distribution indicates a
price increase for electricity. The carbon avoidance cost in Qinghai is the lowest,

38
at 31 £/t, and the highest in Tianjin, at 47.7 £/t. The carbon avoidance cost has
a highly significant effect on production costs, which contribute about 33% of
total production costs.

(Bartels, Pate, and Olson, 2010) They have conducted several studies, and
they found a coal gasification plant has a capital cost of 691 M£ and a plant
capacity of 276,900 kg/day H2. The coal price is 1.29 £/GJ, the production cost
is 2.06 £/kg H2.

Table 8. Comparison between CG plant and SMR plant.

Production Capital cost Feedstock Production Reference


capacity cost cost
SMR plant 215,700 kg 343.3 11.45 £/GJ 2 £/kg H2 (Bartels,
with CCS H2/day Million £ Pate, and
Olson,
2010)
SMR plant 215,700 kg 192.25 9 £/GJ. 1.42 £/kg H2 (Collodi
no CCS H2/day Million £ et al.,
2017)
CG plant 770,700 kg 1238 M£ 1.72 £/GJ 1.4 (Bartels,
with CCS H2/day Pate, and
Olson,
2010)
CG plant 276,900 691 M£ 1.29 £/GJ 2.06 £/kg H2 (Bartels,
with CCS kg/day Pate, and
Olson,
2010)
CG plant 246,260 kg 194.3 M£ 2.14 £/GJ 1.06 £/kg (Li et al.,
(China) no H2/day H2 2022)
CCS
CG plant 246,260 kg 247 M£ 2.14 £/GJ 1.58 £/kg (Li et al.,
with CCS H2/day H2 2022)
(China)

39
The above table shows the effect of the high capital cost of gasification
compared with SMR, but it can benefit from low coal prices. It is also shown that
a larger scale is more economic than a smaller scale. This is due to the
complexity of coal gasification plants, which make less feasible in a small scale.

The findings from these studies highlight the variability in hydrogen production
costs, largely influenced by regional energy prices, feedstock prices, and the
inclusion or exclusion of carbon capture and storage (CCS). It is primarily
affected by the high capital cost of the coal gasification plant. However, while
coal gasification plants have high capital costs due to their complexity, they can
be economical with low coal prices.

4.2.3 Biomass gasification:


Biomass gasification plant capacity of 453,600 kg/day, hydrogen production
cost for the FB gasifier without CCS is 3.04 £/kg H2 with a capital cost of 634
million pounds and 3.43 £/kg H2 with CCS with a capital cost of 836 million
pounds. For the EF gasifier, the hydrogen cost is 3.33 £/kg H2 without carbon
capture with a capital cost of 1206 million £, and 3.44 £/kg H2 with carbon
capture with a capital cost of 1315 million £, the Canadian pine wood biomass
price is 5.1 £/GJ (Salkuyeh, Saville and MacLean, 2018).

Study by Lv et al., (2008) investigated the potential for hydrogen production


from biomass residues in China. The capital cost for the plant was reported to
be 419,300 £, with the feedstock cost at 1.94 £/GJ. With a capacity to produce
517 kg of hydrogen per day and a feedstock feeding rate of 6,400 kg/day, the
plant's production cost was calculated to be 1.93 £/kg H2.

(Bartels, Pate, and Olson, 2010) They have conducted studies about biomass
gasification. A plant with a biomass gasification capacity of 139,700 kg H2/day,
the capital cost of 189 million £, feedstock costs of 3.17 £/GJ, and production
costs of 2.24 £/kg H2. The second study plant with capacity of 194,141 kg
H2/day, a capital cost of 272.4 Million £, feedstock costs of 2.9 £/GJ, the
production cost is 1.62 £/kg H2.

40
Table 9. Summarize biomass gasification plants.

Plant 453,600 453,600 517 139,700 194,141


capacity kg/day (FB kg/day (EF kg/day kg/day kg/day
gasifier) gasifier)
Capital 634 million 1206 419,300 189 272.4
cost £ million £ £ Million £ Million £
without
CCS
Capital 836 million 1315
cost with £ million £
CCS
Feed 5.1 £/GJ 5.1 £/GJ 1.94 3.17 2.9 £/GJ
stock cost £/GJ £/GJ
Production 3.04 £/kg 3.33 £/kg 1.93 2.24 1.62
cost H2 H2 £/kg £/kg H2 £/kg H2
without H2.
CCS
Production 3.43 £/kg 3.44 £/kg
cost with H2 H2
CCS
Reference (Salkuyeh, (Salkuyeh, Lv et (Bartels, (Bartels,
Saville Saville al., Pate, Pate,
and and (2008) and and
MacLean, MacLean, Olson, Olson,
2018). 2018). 2010) 2010)

The primary obstacle for commercializing biomass-derived hydrogen is the high


cost of biomass. Economic assessments indicate that, for the hydrogen
production to be financially viable, the biomass price should be lower 3 £/GJ.
This perspective is corroborated by Salkuyeh, Saville, and MacLean (2018).

From the above studies and data, it can be seen that the cost of hydrogen
production from biomass gasification varies significantly depending on the
feedstock cost, gasifier type, and the inclusion of carbon capture and storage

41
(CCS). Additionally, the scale of the operation appears to have an impact on the
production cost.

4.2.4 Water splitting:


4.2.4.1 Electrolysis:
Electricity costs have a significant impact on the cost of hydrogen production
from electrolysis, which can make up to 85% of the overall cost of hydrogen
production in water electrolysis technologies. In the case of hydrogen
generation from wind energy, Due to the intermittent nature of wind energy, the
plant's capacity plays a crucial role in cost estimation (El-Emam and Özcan,
2019).

Kuckshinrichs, Ketelaer, and Koj (2017) conducted an economic assessment of


the established H2 production method, alkaline water electrolysis (AWE) with a
production capacity of 2838 kg H2/day and a capital cost of 5.82 M£, across
three European locations: Germany, Austria, and Spain. Their findings revealed
that in Germany, the levelized cost of H2 stands at approximately 4.08 pounds
per kilogram. In contrast, both Austria and Spain registered a slightly elevated
levelized cost for H2, marking an increase of 15 to 18%. The primary reason for
this increase is due to higher electricity expenses in the latter two countries, as
shown in the below table.

Table 10. Hydrogen production cost in different region via grid electrolysis.

Germany Austria Spain


Electricity price 7.16 £ p/kWh 8.74 £ p/kWh 9.5 £ p/kWh
Hydrogen 4.08 £/kg H2 4.7 £/kg H2 4.81 £/kg H2
production cost

According to (Nikolaidis and Poullikkas, 2017), several studies have been


carried out that have demonstrated the clean and sustainable nature of water-
splitting pathways. When the electricity necessary for electrolysis is derived
from fossil fuels, CO2 emissions occur. However, when renewable energy
sources such as solar and wind are utilised, both the high capital costs and low
conversion efficiencies increase production cost, in some cases, surpass £20
/kg for electrolysis.

42
Bartels, Pate, and Olson (2010) undertook evaluations of various technologies.
Their findings indicated that solar photovoltaic electrolysis, with a capital cost of
£6,600 million, produces 354,359 kg of H2 per day. The production cost is £7.18
per kg of H2. On the other hand, wind turbine electrolysis, with a production
capacity of 50,000 kg/day of H2, demands an initial capital investment of £671
million. The cost of producing hydrogen via this method is £7.8 per kg of H2.
These observations explain the high capital cost associated with renewable
sources for hydrogen production.

Table 11. Summarize production cost of different electrolysis plants.

Production Capital cost Electricity Cost Hydrogen


Capacity production
cost
Grid 2838 kg H2/day 5.82 M£, 7.16-9.5 £ 4.08-4.81
electrolysis p/kWh £/kg H2
Photovoltaic 354,359 kg 6,600 million£ 7.18 £/kg H2
electrolysis H2/day
Wind turbine 50,000 kg/day £671 million 7.8 £/kg H2
electrolysis

The primary determinant of hydrogen production costs through electrolysis is


the price of electricity. When sourced from renewable energies, solar or wind,
the costs increase to their highest value due to the significant capital
investments required for wind turbines and solar panel infrastructure.

4.2.4.2 Thermochemical:
Several studies on nuclear hydrogen plants (El-Emam and zcan, 2019) have
determined that the cost of nuclear hydrogen using the S-I cycle for a large
capacity plant (1200 tonne/day) is 2 £/kg, while the cost of hydrogen produced
from HyS-MHR for a daily hydrogen production of 580 tonne is 2.23 £/kg. This
relatively reduced cost is most likely due to the large plant capacity.

The IAEA HEEP software was used by El-Emam and Khamis, (2017) for
different nuclear hydrogen generation schemes, PEM electrolysis is integrated
with Advanced Pressurised Water Reactor (APWR), HTGR is used to operate a
High-Temperature Steam Electrolysis (HTSE)-based hydrogen plant, and a S–I
43
plant is evaluated based on various reactor technologies. They reported that the
Gas Turbine High Temperature Reactor (GT-HTR) with S–I plant has the lowest
production cost among S–I cycles at 2.48 pounds per kilogramme of hydrogen
(H2). PEM integrated with nuclear (APWR) has the lowest production cost per
kilogramme of hydrogen at £3.58 and the highest capacity per day at 1,382
tonnes. And High Temperature Gas Cooled Reactor (HTGR) integration with
(HTSE) is the lowest production cost at 2.26 £/kg H2 and lower than integration
with S-I cycle at 2.65 £/kg H2 with the same 345.6 tonne-per-day capacity.

Table 12. Various nuclear hydrogen production strategies El-Emam and


Khamis, (2017).

Hydrogen Nuclear Production Production


production unit reactor type capacity cost
S–I cycle Nuclear 345.6 2.65 £/kg H2
(HTGR) tonne/day
S–I cycle Nuclear (GT- 50.1 tonne/day 2.48 £/kg H2
HTR)
S–I cycle Nuclear (HTR- 117.5 3.81 £/kg H2
PM) tonne/day
PEM Nuclear 345.6 5.5 £/kg H2
(APWR) tonne/day
PEM Nuclear 691.2 4.17 £/kg H2
(APWR) tonne/day
PEM Nuclear 1382 3.58 £/kg H2
(APWR) tonne/day
HTSE Nuclear 345.6 2.26 £/kg H2
(HTGR) tonne/day

(Mehrpooya and Habibi, 2020) They found that large-scale hydrogen production
can reduce capital costs and improve the economics of thermochemical cycles,
making them more competitive with other hydrogen production methods such
as direct electrolysis and steam methane reforming (SMR). They noted that
solar energy is a promising power source for these thermochemical cycles, its
high capital cost currently leads to an increase in the total cost of hydrogen
production. However, they also suggested that anticipated advancements in
44
technology, resulting in cheaper and more efficient processes, are expected to
lower these production costs in the future.

(Orhan, Dincer and Rosen, 2010) they also have observed that Cu-Cl plant
become more economic with larger capacity. They conducted several studies, a
plant with 200 tons/day production cost was 2.25 £/kg, and a plant with 2
tons/day, the production cost was 3.93 £/kg. below table presents these studies.

Table 13. Cost of hydrogen production for Cu–Cl thermochemical plants of


varying hydrogen production capacities (Orhan, Dincer and Rosen, 2010).

H2 production capacity of Cu–Cl plant 2 10 50 200


(tons/day)
Capital cost of plant (£/GJ) 14.9 8.68 4.95 3.05
Capital cost of storage (£/GJ) 0.56 0.56 0.56 0.56
Energy cost (£/GJ) 7.14 7.14 7.14 7.14
Distribution cost (£/GJ) 5.18 5.18 5.18 5.18
Total (£/GJ) 27.75 21.55 17.83 15.91
Total (£/kg) 3.93 3.05 2.52 2.25

The hydrogen production cost from nuclear is derived by plants scale, the larger
scale is more economic. Integrating different reactor types with different
hydrogen production units shows various production costs.

4.3 Uncertainty:
As mentioned before, the economic performance is affected by input
parameters. The uncertainty discusses the variation of input parameter cost.
The prices of natural gas vary from place to place; they are lower in Russia,
which is a producer, and higher in importer countries, so the cost of steam
methane reforming is affected by region prices Fig. 11 below shows hydrogen
costs in different regions.

45
2.5

GBP/kg 1.5

0.5

0
no CCUS with no CCUS with no CCUS with no CCUS with no CCUS with
CCUS CCUS CCUS CCUS CCUS

United States Europe Russia China Middles East

CAPEX OPEX Natural gas

Fig. 11. Levelized hydrogen production from natural gas regionally (IEA,
Hydrogen production costs using natural gas in selected regions, 2018, IEA,

Paris).

The above figure shows that production costs are lowest in countries that
produce natural gas.

Coal gasification is affected slightly by coal prices; China has abundant coal
mines, which have many coal gasification plants and produce hydrogen at
economic costs.

In electrolysis, the price of electricity has the greatest impact on the cost of
hydrogen production. The price of electricity can vary significantly from region to
region, depending on factors like the source of electricity, governmental policies,
and local taxation.

The price of feedstock has a significant impact on biomass gasification costs.


Waste food, due to its abundance and low cost, can be a good source of
feedstock, but its availability can vary depending on the region. (Bartels, Pate,
and Olson, 2010) have collected data about biomass gasification models and
found the biomass prices range from 1.04 to 3.16 £/GJ.

Nuclear thermochemical processes can provide competitive costs for hydrogen


production. However, these costs can be sensitive to plant scaling. As
mentioned, larger-scale plants are typically more economical.
46
The cost of hydrogen production can greatly depend on the method of
production and the region in which the production is taking place. Local factors,
such as the cost of inputs (natural gas, coal, biomass, electricity, etc.) and the
scale of the production plant, can significantly influence the overall production
cost.

47
5 H2A Model
The National Renewable Energy Laboratory (NREL) created the H2A model,
which is a standardised analysis methodology for calculating and comparing the
cost of producing hydrogen using various technologies and pathways. The
"H2A" name stands for hydrogen analysis. The H2A model is designed to allow
for flexible, transparent, and consistent comparisons of the costs of different
hydrogen production technologies. It's a spreadsheet-based model that allows
researchers to examine the impact of different factors on hydrogen production
costs, such as capital costs, operating costs, efficiency, feedstock costs, and
other parameters.

5.1 Natural gas with CO2 sequestration model:


Steam Methane Reforming (SMR) with Carbon Capture and Storage (CCS)
using the H2A model The daily hydrogen production capacity is 483,000 kg/day.
The natural gas price was established at £3.8/MMBtu. The capital cost and
operating cost were reported to be £783 million and £158 million, respectively.
The production cost is £1.82/kg H2.

The model demonstrates a feedstock energy input of 0.182 GJ/kg H2 and a


hydrogen energy output of 0.12 GJ/kg H2, resulting in an efficiency of 65.9%.

In terms of cost contribution, feedstock costs are most significant, constituting


52.6% of the total, as shown in the table below:

Table 14. Cost contribution of SMR model.

Specific Item Cost Calculation

Cost Component Cost Contribution Percentage of H2


(£/kg) Cost
Capital Costs £0.38 20.8%
Decommissioning Costs £0.00 0.1%
Fixed O&M £0.205 11.1%
Feedstock Costs £0.96 52.6%
Other Raw Material Costs £0.039 2.0%
Byproduct Credits £0.00 0.0%

48
Other Variable Costs £0.245 13.3%
(including utilities)
Total £1.82

The CO2 transport and sequestration cost, an additional expense incurred in


the CCS process, is calculated to be £22.3 per tonne of CO2.

The environmental impact is represented by emissions; the total well-to-pump


emissions are 3.042 kg CO2 equivalent per kg of H2.

5.2 Coal gasification with CCS model:


Coal gasification with carbon capture and storage (CCS) using the H2A model.
The daily hydrogen production capacity is 660,000 kg/day. The capital cost and
operating cost were reported to be £2028 million and £184 million, respectively.
The production cost is £2.86/kg H2.

The model demonstrates a feedstock energy input of 0.202 GJ/kg H2 and a


hydrogen energy output of 0.12 GJ/kg H2, resulting in an efficiency of 59.5%.

In terms of cost contribution, capital costs are most significant, constituting


40.9% of the total, as shown in the table below:

Table 15. Cost contribution of Coal gasification model.

Specific Item Cost Calculation

Cost Component Cost Contribution Percentage of H2


(£/kg) Cost
Capital Costs £1.167 40.9%
Decommissioning Costs £0.01 0.2%
Fixed O&M £0.81 28.6%
Feedstock Costs £0.5 17.3%
Other Raw Material Costs £0.039 1.4%

Byproduct Credits £0.00 0.0%


Other Variable Costs £0.33 11.6%
(including utilities)

49
Total £2.86

The CO2 transport and sequestration cost, an additional expense incurred in


the CCS process, is calculated to be £14.7 per tonne of CO2.

The environmental impact is represented by emissions, the total well-to-pump


emissions are 3.4 kg CO2 equivalent per kg of H2.

5.3 Biomass gasification model no CCS:


In the biomass gasification model using woody biomass, the production
capacity is 155,000 kg/day. The capital costs and operating costs were reported
to be 141.7 million £ and 69 million £ respectively. The production cost is 2.36
£/kg H2. Feedstock energy input is 0.264 GJ/kg H2, hydrogen energy output is
0.12 GJ/kg H2, and efficiency is 40%.

In terms of cost contribution, feedstock costs are the most significant,


constituting 56.9% of total costs, as shown in the below table.

Table 16. Cost contribution of biomass gasification model.

Specific Item Cost Calculation

Cost Component Cost Contribution Percentage of H2


(£/kg) Cost
Capital Costs £0.31 13.3%
Decommissioning Costs £0.00 0.1%
Fixed O&M £0.254 10.7%
Feedstock Costs £1.34 56.9%
Other Raw Material Costs £0.20 8.8%

Byproduct Credits £0.00 0.0%


Other Variable Costs £0.24 10.2%
(including utilities)
Total £2.36

50
The environmental impact is represented by emissions, the total well-to-pump
emissions are 27.5 kg CO2-eq/kg H2.

5.4 Grid electrolysis:


A grid electrolysis model with a daily hydrogen production capacity of 56,500 kg
uncovers important insights into its financial and operational characteristics. The
total capital cost and operating cost for this model are estimated at £58.8 million
and £76 million, respectively, leading to a hydrogen production cost of £4.7/kg
H2. The process is marked by an efficiency of 60%, utilizing an electricity
energy input of 0.2 GJ/kg and achieving a hydrogen energy output of 0.12
GJ/kg.

In terms of cost contribution, the electricity is the most significant, constituting


about 85% as shown In below table.

Table 17. Cost contribution of grid electrolysis model.

Specific Item Cost Calculation

Cost Component Cost Contribution Percentage of H2


(£/kg) Cost
Capital Costs £0.39 8.4%
Decommissioning Costs £0.00 0.0%
Fixed O&M £0.235 5.1%
Feedstock Costs £0.00 0.0%
Other Raw Material Costs £0.00 0.0%

Byproduct Credits £0.00 0.0%


Other Variable Costs £4.1 86.5%
(including utilities&
electricity)
Total £4.7

Electrolysis process itself does not emit carbon, the associated carbon
emissions stem from the energy source, estimated at 29 kg CO2-eq/kg H2.

51
The analysis of various hydrogen production models reveals distinct
characteristics and potential applications, each subject to specific constraints
and advantages. The Steam Methane Reforming (SMR) model, with a capacity
ranging from 241,500 to 966,000 kg/day, offers the lowest production cost
(£1.82) and is highly sensitive to natural gas prices. The coal gasification model,
with a capacity between 330,000 and 1,320,000 kg/day, necessitates high
capital investment due to its complexity but remains a favourable option in
regions with abundant coal resources, such as China and the U.S. The biomass
gasification model, limited to 20,000–200,000 kg/day, leverages renewable
resources like woody residue and crops, yet faces challenges related to
biomass availability. Lastly, the electrolysis model, also with a capacity of
20,000–200,000 kg/day, produces green hydrogen with minimal emissions.
Though these emissions are tied to the energy source, this model has the
highest production cost. Environmental impact results show SMR with CCS,
coal gasification with CCS, biomass gasification without CCS, and grid
electrolysis are 3.042, 3.4, 27.5, and 29 kg CO2-eq of carbon emissions,
respectively. SMR has the lowest; coal gasification comes next due to
employing CCS. As shown, biomass gasification without CCS produces high
emissions; the electrolysis process has very low emissions, but emissions come
from energy sources. The cost contributions for SMR, coal gasification, biomass
gasification, and electrolysis are natural gas (52.6%), capital costs (40.9%),
feedstock costs (56.9%), and electricity costs (85%), respectively. The
comparative analysis of these models underscores the intricate interplay of
factors such as resource availability, capital requirements, environmental
impact, and regional considerations in the pursuit of efficient and sustainable
hydrogen production solutions.

52
6 Life Cycle Assessment
Life Cycle Assessment (LCA) is a technique used to evaluate the environmental
aspects and potential impacts of a product. LCA will be conducted on various
hydrogen production technologies. This analysis will include renewable
methods like water electrolysis and biomass gasification, as well as non-
renewable methods like natural gas steam reforming and coal gasification. By
conducting an LCA, the study aims to provide insights into the environmental
impacts of these technologies, such as resource consumption, waste
generation, and carbon emissions. Furthermore, the LCA will also encompass
the carbon capture cost related to each technology, where applicable. This will
ensure a more thorough comprehension of the global environmental effects
associated with each hydrogen production technology.

6.1 Steam Methane Reforming:


A plant size was assumed, processing around 134,820 kg of H2 per day at full
operational capacity. It was noted that such a system necessitates a substantial
electricity input, around 153.311 MJ per day. The report tabulates various air
emissions and details the proportion of each attributable to various subsystems,
such as construction and decommissioning, natural gas production and
transport, and power generation and plant operation. According to Spath and
Mann (2000), carbon dioxide, methane, non-methane hydrocarbons (NMHCs),
NOx, SOx, CO, particulates, benzene, and N2O are the most significant
emissions. The emission is displayed in the tables below.

Table 18. Average air emission.

Air System
Emission total
g/kg of
H2
-Benzene 1.4
-Carbon 10,620
Dioxide
-Carbon 5.7
Monoxide
-Methane 59.8

53
-Nitrogen 12.3
Oxide
-Nitrous 0.04
Oxide
-Non- 16.8
methane
Hydrocarbon
-Particulates 2
-Sulphur 9.5
Oxides

Table 19. Greenhouse Gases Emissions and Global Warming Potential.

Gas GWP value ( g


CO2-
equivalent/kg of
H2)

CO2 10,621

CH4 1256

N2O 11

GWP total 11,888

Assessing the global warming potential (GWP) of the system requires the
quantification of total greenhouse gas emissions, even though CO2 is the
primary greenhouse gas and accounts for the majority of emissions from the
system under evaluation. Considering that the GWP is a cumulative effect of
CO2, CH4, and N2O emissions, this is crucial in this respect. Remarkably, the
warming capacities of CH4 and N2O are 21 and 310 times greater than those of
CO2. Natural gas was the most extensively utilised resource, comprising 94.5%
of the total resources on a weight basis. This was then followed by coal (4,1%),
iron (including ore and waste, 0.6%), limestone (0.4%), and gasoline (0.4%).

54
Iron and limestone were utilised predominantly in the construction of the power
plant and pipeline (Spath and Mann, 2000). Table 20 presents average
consumption.

Table 20. Average resource consumption.

Resource total % of
(g/kg Total in
H2) this
table

Coal 159.2 4.10%

Iron 10.3 0.30%

Iron scrap 11.2 0.30%

Limestone 16 0.40%

Natural 3,642.30 94.50%


gas
Oil 16.4 0.40%

Water consumption was 19.8 litres per kilogramme of hydrogen, with 95.2% of
the water consumed by the hydrogen plant for reforming and shift reactions, as
well as steam production. Regarding water pollutants, the total volume was
calculated to be 0.19 g/kg of H2, with hydrocarbons (60%) and dissolved matter
(29%) being the most prevalent contaminants. As for waste, the system
generated primarily non-hazardous waste at a rate of 201,6 g/kg of hydrogen
produced. The vast majority (72.3%) of this waste originated from the
production and distribution of natural gas. Further analysis revealed that
pipeline transportation accounted for fifty percent of the total system waste, with
natural gas extraction accounting for twenty-two percent (Spath and Mann,
2000).

6.2 Coal Gasification:


A coal gasification plant, situated approximately 2000 km from Toronto in
Stellarton, Nova Scotia, have been disclosed by the Coal Association in

55
Canada. This is the additional distance between the Nanticoke coal-fired power
plant and Toronto. The coal gasification plant, which does not include CO2
capture, is characterized by a daily hydrogen production of 284 tonnes, a daily
coal feed of 2268 tonnes, and a net power output of 38 MW. A surface mine
provides all of the plant's coal requirements. (Cetinkaya, Dincer, & Naterer,
2012) Table 14 displays a negative value for electric power production,
indicating the plant's electricity export to the grid. The life cycle energy
consumption per kg of hydrogen product is estimated 213.8 MJ (Li and Cheng,
2020).

Table 21. Coal gasification plant air emissions (Cetinkaya, Dincer and Naterer,
2012).

Process CO2 equivalent emission [g CO2-e/kg


H2]
Surface mining 101.64
Transport of coal 162.34
Mine development 60.50
Manufacture/construction of the plant 48.40
Electric power production -447.70
Coal consumption 11374.00
Total 11299.18

6.3 Biomass gasification:


Biomass gasification plant, 500 tons of wheat straw (WS) are utilized daily, with
an optimal gasification temperature of 700°C. Operating under standard
atmospheric pressure, the pyrolizer, combustor, gasifier, and reformer achieve
temperatures of 600°C, 900°C, and 700°C, respectively. The Pressure Swing
Adsorption (PSA) component, showcasing a remarkable 97% separation
efficiency, runs at 40°C and 1.6 MPa. The system incorporates a two-stage
compressor characterized by a pressure ratio of 4 for a single stage. The
isentropic efficiency of this compression phase is 0.8, while its mechanical
efficiency stands at 0.95. The PSA's prowess in hydrogen separation remains
consistent at 97%. The primary steam in the affiliated power facility operates at
a temperature and pressure of 540°C and 16.7 MPa, respectively. As
56
documented by Li et al., 2020, the overall system efficiency measures 58.8%.
Subsequent tables present the consumption of resources and associated air
emissions.

Table 22. Resource consumption of gasification process (Li et al., 2020).

Resource
Consumption
Biomass 13822.2 (g/kg-H2)

Fertilizer and 489.3 (g/kg-H2)


pesticide

Diesel 238.5 (g/kg-H2)

Electricity 2459.4 (Wh/kg-H2)

Table 23. Air emission from gasification process (Li et al., 2020).

Emission g/kg-H2

O2 2367.8

N2 32437.1

H2O 2796

CO2 6971.7

CH4 11.3

N2O 2.69

NOx 17.9

SO2 11

CO 7.84

HC 2.53

PM10 2.05

57
COD 2.76

Solid 261.6
waste
TN 33.8

NH3–N 0.88

Pesticide 6.2

6.4 Water splitting:


6.4.1 Wind electrolysis:
According to (Bhandari, Trudewind, and Zapp, 2014), the system is comprised
of three 50-kW wind turbines. This system is coupled with an electrolyser
capable of producing 64 kg of hydrogen per day with a conversion efficiency of
approximately 85 percent. Produced hydrogen is compressed, stored, and then
distributed to fueling stations. In this process chain, fossil fuels, metals, and
minerals are crucial inputs. Iron, which is used extensively in the construction of
wind turbines and hydrogen storage vessels, accounts for 37.4% of the utilised
resources. A considerable quantity of limestone, amounting to 35.5% of the
main resources, is used to create the turbines' concrete foundations. Coal,
which is predominantly employed in the production of steel, iron, and concrete,
accounts for 20.8% of the deployed resources. Following these are oil and
natural gas, which are used primarily in the production of wind turbines and
account for 4.7% and 1.6% of resource consumption, respectively. Water is
another essential resource, utilised not only during electrolyzer operation but
also in upstream processes at a rate of approximately 26,7 litres per
kilogramme of hydrogen. Approximately 45% of this water is utilised by the
electrolyser, 38% by the production of wind turbines, and 17% by the
manufacturing of hydrogen storage vessels. The cumulative resource
consumption values for this process are summarized in Table 24.

58
Table 24. Cumulative consumption of resources in the wind electrolysis system
(Bhandari, Trudewind and Zapp, 2014).

Resource Total (g/kgH2) Wind turbines Electrolysis Storage (%)


(%) (%)
Coal 214.7 68 5 27
Iron (Fe, ore) 212.2 64 6 30
Iron scrap 174.2 53 8 39
Limestone 366.6 96 1 3
Natural gas 16.2 72 15 13
Oil 48.3 76 13 11

The average energy consumption per kilogramme of hydrogen was 9.1 MJ. A
substantial portion of this energy consumption, specifically 72.6%, was
attributable to the wind turbine manufacturing process. The amounts of energy
consumed by electrolysis and storage were significantly lower, at 4.8% and
31.1%, respectively. In terms of emissions, CO2 was the major contributor,
accounting for approximately 95% by weight. The majority of these air
emissions were produced during the manufacturing and installation of wind
turbines. Table 25 provides a comprehensive summary of the air emission
values for the system.

Table 25. Cumulative air emissions from electrolysis driven by the wind
(Bhandari, Trudewind and Zapp, 2014).

Air emission Total (g/kgH2) Wind tur-bines Electrolysis Storage (%)


(%) (%)
Carbon 950 78 4 18
dioxide
Carbon 0.9 80 4 16
monoxide
Methane 0.3 92 3 5
Nitrogen 4.7 46 47 7
oxides

59
Nitrous oxides 0.05 67 6 27
Non-methane 4.4 63 7 30
hydrocarbons
Particulates 28.7 94 1 5
Sulfur dioxide 6.1 62 26 12

Various studies have examined the global warming potential (GWP) associated
with wind-based electrolysis. Remarkably, these studies have all reported a
consistent GWP value of 0.97 kg of CO2 equivalent per kilogram of hydrogen.
This consistency across different research efforts underscores the
environmental efficiency of wind-based electrolysis in hydrogen production
(Bhandari, Trudewind, and Zapp, 2014).

6.4.2 Electrolysis integrated with PV power:


According to (Cetinkaya, Dincer, and Naterer, 2012), a building-integrated
photovoltaic (PV) system with a capacity of 8 kW and an area of 160 square
metres is projected to have a lifespan of 30 years. The PV panels, in this
context, are positioned roughly 32 kilometres from the heart of Toronto. The
total travel distance is approximated to be 600 kilometres. This distance
involves the journey of the panels from their place of manufacture to Sarnia and
subsequently to the city centre of Toronto. In Table 26 are detailed the energy
equivalents for the solar power plant and the corresponding carbon dioxide
equivalent emissions for each step of the process. This data provides a
comprehensive analysis of the environmental impact of this PV system over its
operational lifetime.

Table 26. Carbon dioxide equivalent emissions and solar energy plant energy
equivalents (Cetinkaya, Dincer and Naterer, 2012).

Processes Energy equivalent CO2 equivalent emissions


[kJ/kg H2] [g CO2-e/kg H2]
Materials and manufacturing 25,550.48 1519.53
of PV modules
Transportation 602.53 461.36
Inverters 830.91 110.93

60
Wiring 602.41 60.24
Installation 2679.68 37.18
Operation and maintenance 2285.00 161.20
Decommissioning and 893.23 61.70
disposal
Total 33,444.24 2412.13

6.4.3 High temperature electrolysis for hydrogen production via nuclear


energy:
The system consists of a high-temperature nuclear plant integrated with high-
temperature electrolysis. This electrolysis consumes both electricity and thermal
energy from nuclear plant. The nuclear plant has a thermal capacity of 600 MW.
The energy is divided to provide thermal energy and electric power to the
electrolysis system, and the residue power is directly sent to the national grid.
The primary components of the electrolysis module encompass heat
exchangers, water purification mechanisms, hydrogen isolation units, water
supply systems, circulatory pumps, and a planar electrolysis stack assembly.
For producing 1 kg of hydrogen, the thermal requirement is 235 MJ, 200 MJ is
electric energy, and 35 MJ is heat energy. The hydrogen production rate from
600 MW (500 MW to hydrogen production) is 184,800 kg per day (Utgikar and
Thiesen, 2006).

The cumulative emissions from this system are the sum of the nuclear plant and
the HTE production unit. For each 1 kg of produced hydrogen, the emissions
are 450g of CO2, 730 mg of SO2, and 1600 mg of NOx. The primary source of
emissions is the lifecycle stage of fuel extraction, processing, and
transportation, which represents 63% of total emissions. Other stages such as
materials procurement, construction, operational activities, and waste
management contribute 5%, 8%, 15%, and 9%, respectively (Utgikar and
Thiesen, 2006).

The emissions associated with manufacturing solid oxide fuel cells, including
materials required for construction and equipment for the balance of the unit,
are delineated as follows: CO2 levels ranging from 45 to 25 g, SO2 levels
ranging from 0.5 to 1.1 g, and NOx levels ranging from 0.2 to 2.7 g. The

61
extraction and processing of metals like iron, nickel, and chromium emerge as
major contributors, accounting for 40% of the total carbon footprint and half of
the acidification effects. Nuclear-HTE plant emissions are measured to be 700g
CO2, 1.8 SO2, and 4.3 NOx for each 1 kg of H2. NOx possesses a global
warming potential (GWP) 310 times that of CO2. Consequently, for this system,
the GWP is tabulated at 2 kg CO2-equivalent per kilogram (Utgikar and
Thiesen, 2006).

6.5 Environmental impact assessment:


Environmental Impact Assessment (EIA) is a systematic process that identifies
and evaluates the potential effects of proposed projects or policies on the
natural environment. It is an essential tool in environmental management and is
used to predict the consequences of actions before they are carried out,
allowing for the avoidance or mitigation of harmful impacts and enhancing
beneficial effects. Common indicators utilised in the evaluation of life cycle
environmental consequences include global warming potential (GWP),
acidification potential (AP), eutrophication potential (EP), abiotic depletion
potential (ADP), and human toxicity potential (HTP). According to Ji and Wang
(2021), the comprehensive measurement of greenhouse gas emissions
associated with the entire process is a crucial component that greatly affects the
environmental implications of hydrogen generation. The inventory data utilised
for the assessment is in Table 27.

Table 27. GWP and AP of different hydrogen production methods, represent


data as average values (Ji and Wang, 2021).

Production method Average GWP (kg CO2 Average AP (g SO2-eq/kg


eq/ kg H2) H2)
SMR 11.98 15.2
SMR with CCS 3.7
CG 22.99 59.7
CG with CCS 4.87
Biomass gasification 3.54 22.5
Wind based electrolysis 1.08 4.3
Solar based electrolysis 1.82 6.1

62
The GWP and AP reported from life cycle assessment are in table below:

Table 28. GWP and AP reported from life cycle assessment.

Production method GWP kg CO2-eq/kg H2 AP g SO2-eq/kg H2


SMR no CCS 11.888 18.11
Coal gasification no 11.299
CCS
Biomass gasification no 8 23.53
CCS (Wheat straw)
Wind electrolysis 0.97 9.39
PV power electrolysis 2.412
HTE based on nuclear 2 1.8

Fossil fuel-derived processes, such as coal gasification (CG) and the reforming
of natural gas, are observed to have the most detrimental environmental
impacts due to their extraordinarily high GWP and AP values. Specifically, CG,
with its GWP of 22.99 kg CO2 equivalent, ranks second among all hydrogen
production methods. Implementing Carbon Capture and Storage (CCS)
technology in the fossil fuel hydrogen production process could lead to
significant reductions in greenhouse gases. It's projected that the use of CCS
could reduce the GWP of CG by between 71.7% and 81.7%. For steam
methane reforming (SMR), a decrease in GWP of around 69.1% is expected on
average. The actual impact depends on the effectiveness of CO2 sequestration.
The primary energy consumption calculations of 153 MJ for steam methane
reforming (SMR) and 213.8 MJ for coal gasification (CG) suggest that SMR is a
more energy-efficient choice.

Regarding global warming potential (GWP) and acidification potential (AP),


biomass gasification occupies an intermediate position. It has fewer
environmental impacts than methods based on fossil fuels, but more than
thermochemical cycles based on nuclear energy and electrolysis based on
renewable energy sources. Given the current technological advances, biomass
gasification is a viable transitional option for transitioning from fossil fuels to
water as the primary source of hydrogen. According to (Salkuyeh, Saville and

63
MacLean, 2018), biomass gasification plant use Canadian Pine wood. The CO2
emissions from FB gasification without carbon capture is 27.3 kg CO2/kg H2,
and with CCS is 11.4 kg CO2/kg H2. For EF the CO2 emissions without carbon
capture is 21.4 kg CO2/kg H2, and with CCS, is 0 CO2/kg H2. When a
CCS unit is used in conjunction with an EF gasification process, there are
virtually no CO2 emissions. This is because EF gasification achieves high
temperatures and allows for more complete conversion of feedstock into
syngas, with reduced CO2 emissions. In the FB gasification process, only about
60% of CO2 is captured and liquefied. The CO2 emission from biomass is
varied depending on biomass composition.

The global warming potential (GWP) of the electrolysis depends significantly on


the source of electrical energy. If grid electricity is used in accordance with the
present electricity balance, the environmental impact is quite severe, with a
global warming potential (GWP) of 29.21 kg CO2 equivalent and an acidification
potential (AP) of 67.0 g SO2 equivalent. These figures surpass coal gasification
(CG) (22.99 kg CO2 equivalent, 59.7 g SO2 equivalent) and steam methane
reforming (SMR) (11.98 kg CO2 equivalent, 15.4 g SO2 equivalent) (Ji and
Wang, 2021). However, the GWP and AP are considerably reduced if the
electricity is produced using wind, solar, or nuclear energy. This result highlights
the significance of combining renewable electricity sources with electrolysis
systems for hydrogen production.

Thermochemical cycles powered by nuclear energy tend to have a relatively


minor environmental impact. Compared to some other hydrogen production
techniques, the High Temperature Electrolysis (HTE) process using nuclear
energy, as described in the study by Utgikar and Thiesen (2006), has relatively
low emissions. Global Warming Potential (GWP) is 2 kg CO2-eq/kg H2, and
Acidification Potential (AP) is 1.8g SO2-eq/kg H2.

According to a study by Karaca, Dincer, and Gu (2020), high temperature


electrolysis (HTE) employing nuclear power exhibited the least environmental
impacts among the various hydrogen production methods studied. This
superiority of HTE was markedly noticeable in the realms of global warming
potential (GWP) and acidification potential (AP), where it surpassed alternatives

64
such as the five-step Cu–Cl cycle. In particular, the GWP for HTE was
calculated at 0.477 kg CO2-eq per kg of hydrogen, while for the five-step Cu–Cl
cycle, it was significantly higher at 1.346 kg CO2 eq per kg of hydrogen.

6.6 Carbon capture storage:


Carbon Capture and Storage (CCS) represents a technological approach that
can sequester approximately 90% of the carbon dioxide (CO2) emissions
stemming from fossil fuel consumption in power production and various
industrial activities. This ensures that CO2 is prevented from being released into
the atmosphere. Subsequent to its capture, the CO2 is conveyed and securely
stored beneath the Earth's surface.

In hydrogen production methods that depend on fossil fuels, CCS is important


to mitigate CO2 emissions. These methods of producing hydrogen from fossil
fuels are steam methane reforming (SMR) and coal gasification. Both of these
processes release significant amounts of CO2 as byproducts. By incorporating
CCS technology, these emissions can be significantly reduced, contributing to a
lower overall carbon footprint for hydrogen production. It is also used in biomass
gasification to capture CO2. It's important to note that the efficiency of CCS
technology is dependent on the effectiveness of CO2 sequestration.

6.6.1 Steam methane reforming:


The Steam Methane Reforming (SMR) process does indeed produce
substantial greenhouse gas (GHG) emissions. The reported life cycle emissions
(LCE) were 11.98 kg of CO2 per kg of H2 in scenarios without carbon capture.
However, when carbon capture technology is employed, the emissions
significantly decrease to a range of 3.7 kg of CO2 per kg of H2. Approximately
60% arise from the reforming process itself, where methane (CH4) is reacted
with water (H2O) to produce hydrogen (H2) and carbon dioxide (CO2). The
remaining 40% of CO2 emissions are associated with the heat and power
sources that drive the SMR process. In order to capture CO2, the amine solvent
monoethanolamide (MEA) is often used. This substance can absorb up to 90%
of the CO2 present, which is then typically compressed and stored
underground. This application of Carbon Capture and Storage (CCS)
technology allows for a significant reduction in the greenhouse gas emissions

65
associated with the SMR process, making it a more sustainable option for
hydrogen production.

Integrating Carbon Capture Storage (CCS) technology into a Steam Methane


Reforming (SMR) facility indeed adds an extra cost to the hydrogen production
process. Researchers have been examining this additional cost over the past
couple of decades, typically expressing it in terms of carbon avoidance cost
(CAC), measured in dollars per tonne of CO2 ($/t CO2). The reported CAC for
SMR with CCS, along with the year of the published study, is shown in Fig. 13
(Parkinson et al., 2019).

Fig. 13. Time series cost of avoided carbon ($2016 t1 CO2) (Parkinson et al.,
2019).

According to the 2017 IEAGHG technical report, a steam methane reforming


(SMR)-based hydrogen plant equipped with carbon capture storage (CCS)
could expect a carbon avoidance cost (CAC) ranging between £58.22 and
£87.33 per tonne of CO2, depending on the level of capture (between 56% and
90%). This study used a figure of £12.45 per tonne of CO2 for the transport and
storage costs of the captured carbon. For a scenario with 90% carbon capture,
the CAC was calculated to be £86.8 per tonne of CO2. With an adjusted
transport and storage cost of £22 per tonne of CO2, the total CAC for equipping
CCS at an SMR plant to achieve 90% capture rises to £92.5 per tonne of CO2.

A thorough study by Collodi et al. (2017) evaluated various steam methane


reforming (SMR) plant configurations. Their findings revealed that the cost of
carbon avoidance varied from £44 to £77.2 per tonne of CO2, a range that was

66
closely correlated with the rate of carbon capture within the system, which
spanned from 53% to 90%. This highlights the influence of the capture rate on
the overall cost-effectiveness of carbon avoidance in the operation of SMR
plants.

Varbanov et al. (2020) assessed a steam methane reforming (SMR) plant


equipped with a carbon capture system operating at a 90% capture rate. Their
findings revealed that the cost associated with carbon capture constituted a
substantial 57% of the total hydrogen production cost, translating to £70 per
tonne of CO2 captured. This demonstrates the significant cost that carbon
capture technologies add to hydrogen production from SMR plants.

6.6.2 Coal gasification


Coal, compared to natural gas, has a lower carbon-to-hydrogen ratio, which
results in higher CO2 emissions in the hydrogen production process. Emissions
range from 4.87 and 22.99 kg of CO2 per kg of H2 with and without CCS,
respectively. integration of carbon capture storage (CCS), capturing 90% of the
emissions.

Li et al. (2022) report that, given China's present electricity prices, the cost of
CO2 capture is relatively low, at approximately 23.63 £/t CO2. The CO2
conveyance system consists of the main pipeline connecting the hydrogen
facility to the storage location, as well as branch pipelines leading to each
injection well. With a main pipeline extending over 200 kilometres and a branch
pipeline spanning 30 kilometres, the projected transportation costs are 7.9
pounds per tonne of carbon dioxide. representing 8.25% of total CCS
expenditures. The avoided cost is 40,8 £/t CO2, which is greater than the sum
of the costs associated with capture, transport, and storage. This is due to the
fact that fewer greenhouse gas emissions were produced.

Regional variations in the avoided costs associated with carbon capture and
storage (CCS) across different provinces in China The differences in these
costs are primarily influenced by electricity prices. The regional distribution
shows an increase from west to east in China. Qinghai, located in the western
part of China, boasts the lowest avoided cost of 30.64 £/ton CO2. In contrast,
provinces in East China, North China, and Northeast China display higher

67
avoided costs. The highest is found in Tianjin, where the avoided cost is 47
£/ton. This regional variance is essential for policymakers and industries to
consider when planning and implementing CCS projects across China. Li et al.
(2022).

6.6.3 Biomass gasification:


Biomass gasification emissions reported from life cycle assessment are 8 kg
CO2/kg of H2 for wheat straw biomass; also reported emissions for biomass
gasification with CCS is 3.54 kg CO2/kg H2. For Canadian pine wood biomass,
the emission is 27.3 kg CO2/kg H2, and employing CCS reduces emissions to
11.4 kg CO2/kg H2 with FB gasifier and EF gasifier it reduce to zero emission.
This presents the importance of carbon capture.

Due to the similarities between the processes of coal and biomass gasification,
the Carbon Avoidance Cost (CAC) determined for coal gasification, which is £40
per tonne of CO2, is also applied as the avoidance cost for biomass gasification
(Parkinson et al., 2019).

Poluzzi, Guandalini, and Romano (2022) have assessed biomass for hydrogen
production plants. Their research revealed that the plant can achieve a carbon
capture efficiency ranging between 64% and 90%. The cost associated with this
carbon capture process was estimated to be within the range of 48-51 £ per
tonne of CO2.

6.6.4 Electrolysis:
The direct emissions associated with the electrolysis process are notably low.
However, this doesn't account for the indirect emissions that arise from the
electricity feedstock and the electrolyser system, which must be included in any
comprehensive assessment. The theoretical minimum energy requirement for
this process is about 39.3 kWh per kg of H2. Nevertheless, real-world
commercial applications generally require a more substantial energy input,
usually ranging from 50 to 60 kWh. This energy requirement can lead to
significant emissions when the electricity comes from a grid that's largely
powered by fossil fuels. In such cases, the total emissions related to the
electrolysis process can even exceed those associated with hydrogen
production via natural gas reforming (Parkinson et al., 2019).

68
7 Result and Discussion
7.1 Production Cost:

Production cost values for differnt methods


25
Production Cosr £/kg H2

20

15

10

0
SMR Coal Gasification Biomass Grid Electrolysis Electrolysis Nclear-based
Gasification based on hydrogen
renewable
sources
Production method

Fig. 13. Hydrogen production costs for different methods collected from
literatures.

Hydrogen sources derived from SMR offer a cost-effective method for producing
hydrogen and are expected to continue to dominate this field. Even though coal
plants have a higher capital expense than SMR plants, their feedstock is less
expensive. In addition, the United States' confirmed coal reserves have the
potential to last for centuries. In contrast, natural gas plants have relatively low
capital costs, but the cost of the feedstock is significantly higher, and proven
reserves are considerably lower than those of coal. Given the relative
affordability of coal, the cost of hydrogen production in coal plants is primarily
based on the plant's initial capital cost. In contrast, the cost of hydrogen
production in natural gas plants is highly dependent on the price of natural gas.

Sensitivity to feedstock cost plays a pivotal role in calculating the profitability of


systems where the feedstock is utilised, along with its impact on the levelized
cost of hydrogen under varying feedstock prices. According to Bartels, Pate and
Olson (2010), the cost of feedstock bears the potential to dictate the most
profitable method for hydrogen production and to influence the global market

69
price of hydrogen. SMR and gasification methods produce a significant amount
of CO2. Employing CCS can capture about 70-90% of emissions, which adds
an additional cost of about 30%.

Biomass presents an attractive cost range for the production of hydrogen.


Concerning biomass, however, is the inadequacy of land to simultaneously
generate the required amount of fuel and satisfy global food demand. If waste
streams, such as municipal solid waste or other organic matter, can be utilised
in these processes, then biomass could substantially contribute to hydrogen
production as a renewable feedstock. The cost of the biomass feedstock is a
crucial factor that can affect the economic viability of these processes (Bartels,
Pate & Olson, 2010).

The price of electricity has a significant impact on the cost of electrolysis to


produce hydrogen. Using renewable sources such as wind to generate
electricity for hydrogen production is highly costly due to the high capital costs
of wind turbines and solar panels.

Nuclear-based hydrogen production has the potential to be both competitive


and scalable, largely due to the benefits inherent in nuclear power. It is noted
that plant scale has an influence on production costs.

Levelized Cost of Hydrogen


5
Production Cost £/kg H2

4
3
2
1
0
SMR with CCS CG with CCS Biomass Gasification Grid Electrolysis
no CCS
Production Method

Other Variable Costs (including utilities& electricity)


Other Raw Material Costs
Feedstock
Fixed O&M
Capital costs

Fig. 14. Levelized cost of hydrogen from H2A models.

70
Fig. 14 represents the production cost and cost contribution for each parameter
for each technology in the H2A models. Steam Methane Reforming (SMR)
stands out as the method with the lowest hydrogen production cost at 1.82 £/kg
H2. SMR is primarily dependent on natural gas prices, which contribute 52.6%
to the total cost, and operates at an efficiency of 65.9%. Meanwhile, coal
gasification is associated with a production cost of 2.86 £/kg H2 and is heavily
dependent on capital costs, constituting 40.9% of the total cost. The process
operates at an efficiency of 59.5% and may reap benefits from the low prices of
coal.

In comparison, biomass gasification without carbon capture and storage (CCS)


provides a production cost of 2.36 £/kg H2 and operates at a relatively low
efficiency of 40%. Its total cost is largely sensitive to feedstock prices, which
account for 56.9% of the overall cost. The costliest method is grid electrolysis,
with a production cost of 4.7 £/kg H2. This process is highly sensitive to
electricity prices, which make up about 80% of the total cost, and operates at an
efficiency of 60%.

7.2 Sensitivity:
7.2.1 Steam methane reforming:

SMR Sensitivity
3.5
Production cost £/kg H2

3
2.5
2
1.5
1
0.5
0
5.5 6.25 9 11.45
Natural gas price £/GJ

SMR CCS SMR no CCS

Fig. 15. SMR sensitivity with natural gas prices from collected data.

71
Fig. 15. Representing the sensitivity of hydrogen production costs from natural
gas, it shows increasing production costs with increasing natural gas prices and
also shows the effect of employing CCS.

7.2.2 Coal gasification:

Coal gasification sensitivity


2.5
Production Cost £/kg H2

1.5

0.5

0
276 770
Production Capacity tonne H2/day

Fig. 16. Coal gasification production cost with different scales from collected
data.

The above figure shows that the large scale is more economical, but due to the
complexity of the coal gasification plant, the small scale is less feasible. Even
the larger-scale plant operates with a higher coal price of 1.72 £/GJ than the
smaller-scale plant at 1.29 £/GJ. This shows that plant scale is the biggest
contributor to production costs.

72
7.2.3 Biomass gasification:

Biomass Gasification Sensitivity


3.5
production cost £/kg H2
3
2.5
2
1.5
1
0.5
0
1.94 2.9 3.17 5.1
Biomass price £/GJ

Fig. 17. Biomass gasification sensitivity with biomass price from collected data.

The above figure shows the effect of an increase in biomass prices on hydrogen
production costs. Production costs increase with increased biomass prices. The
biomass price is the main contributor to the production cost.

7.2.4 Electrolysis:

Electrolysis sensitivity
5
Production cost £/kg H2

4.8
4.6
4.4
4.2
4
3.8
3.6
7.16 8.74 9.5
Electricity prices £ p/kWh

Fig. 18. Electrolysis sensitivity from collected data.

The above chart shows increasing production costs with increasing electricity;
electricity costs contribute about 85% of total production costs.

73
Integrate electrolysis with renewable sources (wind turbines, solar panels);
capital costs are the main contributor to the high cost of infrastructure.

7.2.5 Nuclear hydrogen production:

Cu-Cl Plant Sensitivity


4.5
Production cost £/kg H2

4
3.5
3
2.5
2
1.5
1
0.5
0
2 10 50 200
Plant Capacity tonnes/day

Fig. 18. Cu-Cl plant sensitivity.

The above figure shows that the larger scale of a nuclear hydrogen plant is
more economical. Also, the configuration of different reactor types with different
hydrogen plan units has an effect, as discussed in the economic evaluation
chapter.

7.3 Environmental Impact:


From an environmental perspective reported from H2A models, well-to-pump
emissions vary across these technologies. SMR and coal gasification with CCS
emit 3.04 and 3.4 CO2-eq, respectively, demonstrating the significant role of
CCS in reducing environmental impact. Biomass gasification without CCS and
grid electrolysis, on the other hand, have higher emissions of 27.5 and 29 CO2-
eq, respectively. Despite the low emissions associated with the electrolysis
process, the energy source for the process stands out as a significant
contributor to its environmental impact.

Environmental impact assessments for different technologies reported from life


cycle assessments are shown in the below figure.

74
GWP, AP for different production methods
70
CO2-eq kg/kg H2/ SO2-eq g/kg H2 60

50

40

30

20

10

Production method
CO2-eq kg/kg H2 AP g SO2-eq/kg H2

Fig. 19. Environment impact assessment for different production methods GWP
and AP values.

Fig. 19 provides data indicating the environmental impact of various hydrogen


production methods. Well-to-wheels (or life-cycle) CO2-equivalent emissions
per kilogramme of hydrogen produced (CO2-eq kg/kg H2) and acidification
potential (AP), measured in grammes of SO2-equivalents per kilogramme of
hydrogen, are the two measurements.

Steam Methane Reforming (SMR) without Carbon Capture and Storage (CCS)
has CO2-eq emissions of 11.98 kg/kg H2 and an AP of 15.2 g SO2-eq/kg H2.
With CCS, the CO2-eq emissions for SMR reduce significantly to 3.7 kg/kg H2.
Coal gasification (CG) without CCS presents a higher environmental impact
because coal contains more contaminants than natural gas. With CO2-eq
emissions of 22.99 kg/kg H2 and an AP of 59.7 g SO2-eq/kg H2, when CG is
coupled with CCS, the CO2-eq emissions decrease to 4.87 kg/kg H2.

75
For biomass gasification (BG), there are multiple scenarios. Without CCS and
using wheat straw as feedstock, the CO2-eq emissions are 8 kg/kg H2, and the
AP is 23.53 g SO2-eq/kg H2. Using woody biomass (H2A Model) without CCS,
the CO2-eq emissions are significantly higher at 27.5 kg/kg H2, due to
compositional variation. However, BG with CCS shows the lowest CO2-eq
emissions of 3.54 kg/kg H2 among all gasification and reforming methods, and
its AP is 22.5 g SO2-eq/kg H2. Salkuyeh, Saville, and MacLean (2018) have
assessed EF gasifiers with carbon capture, resulting in zero emissions due to
EF gasification, which achieves high temperatures and allows for a more
complete conversion of feedstock into syngas.

It noticed the importance of implementing carbon capture (CCS) to reduce


emissions by up to 90%. This is especially significant in methods such as steam
methane reforming (SMR) and coal gasification (CG), where the use of CCS
drastically reduces their environmental impact. It adds an additional production
cost of about 30%. The reported costs for SMR, CG, and biomass gasification
are 44–92.5, 30.64–47, and 40–51 pounds per tonne of CO2.

Electrolysis, based on renewable energy sources, seems to be the most


environmentally friendly. When using wind power, CO2-eq emissions are only
1.08 kg/kg H2 with an AP of 4.3 g SO2-eq/kg H2, whereas solar power
produces CO2-eq emissions of 1.82 kg/kg H2 with an AP of 6.1 g SO2-eq/kg
H2. However, grid electrolysis presents the highest CO2-eq emissions of 29
kg/kg H2. Due to grid electrolysis's significant emissions from the use of fossil
fuels as an electric source, it is not environmentally friendly. Electrolysis based
on wind, solar, and nuclear energy solves this problem and presents low carbon
emissions.

High-temperature electrolysis has relatively low emissions and is close to


renewable sources. Global Warming Potential (GWP) is 2 kg CO2-eq/kg H2,
and Acidification Potential (AP) is 1.8 g SO2-eq/kg H2.

76
8 Conclusion And Recommendation
In conclusion, natural gas reforming is the most cost-effective method for
hydrogen production, and coal comes next with the benefit of low coal prices,
but both are non-renewable. Biomass gasification is considered a renewable
resource and provides a promising hydrogen production cost, but its limitation is
the availability of biomass feedstock. Producing hydrogen from natural gas and
gasification emits CO2 at 10–25 kg CO2/kg H2, but it can be reduced to 90% by
employing carbon capture facilities. Hydrogen production from electrolysis using
renewable resources is still expensive due to the high capital cost of wind
turbines and solar panels, but it provides very clean hydrogen with very low
emissions. Thermochemical cycles based on nuclear energy offer a competitive
solution, both economically and environmentally.

Methane pyrolysis is not yet commercial, but it will be an alternative to SMR due
to its very low emissions and lower production costs because it does not require
CCS. SMR is preferred to use when natural gas is available at low prices, while
coal gasification is preferred to use when abundant coal is available. Fossil
fuels are non-renewable sources. Wind and solar plants are expensive at this
time due to their high capital costs, so biomass gasification, considered a
renewable source, will be a transition option to renewable resources. Also, it is
important to employ CCS in SMR and gasification plants to mitigate
environmental impact, as that is the goal that led to the choice of hydrogen as
an energy carrier. Hydrogen production from wind turbines and solar panels is
used to balance the supply and demand of energy. There are times when these
sources produce more energy than is needed (like a sunny day with strong
winds) and store it as hydrogen, and times when they produce little to no energy
(like at night or on a calm day). This process is known as "power-to-gas" (P2G).

Hydrogen production based on nuclear integrated with HTE units presents


competitive production costs, large scale, and low emissions. Nuclear energy is
anticipated to be the primary driver of hydrogen production in the future.

With a higher energy content of 120 GJ/tonne, hydrogen is a superior energy


carrier compared to natural gas and coal, which contain 39 GJ/tonne and 29
GJ/tonne, respectively. The significant energy density of hydrogen makes it an
attractive option for various applications, including power generation and
77
transportation. Moreover, when used in a fuel cell, it generates electricity with
water being the only byproduct, making it an environmentally friendly energy
carrier. Hence, hydrogen, particularly when produced from renewable or low-
carbon sources, has the potential to play a crucial role in achieving a
sustainable and low-carbon energy future.

78
References :

Ball, M., & Wietschel, M. (2009, January). The future of hydrogen –


opportunities and challenges. International Journal of Hydrogen Energy,
34(2), 615–627.

Bartels, J.R., Pate, M.B. and Olson, N.K. (2010). An economic survey of
hydrogen production from conventional and alternative energy sources.
International Journal of Hydrogen Energy, [online] 35(16), pp.8371–8384.
doi:https://doi.org/10.1016/j.ijhydene.2010.04.035.
Bhandari, R., Trudewind, C.A. and Zapp, P. (2014). Life cycle assessment
of hydrogen production via electrolysis – a review. Journal of Cleaner
Production, [online] 85, pp.151–163.
doi:https://doi.org/10.1016/j.jclepro.2013.07.048.
Carmo, M., Fritz, D.L., Mergel, J. and Stolten, D. (2013). A comprehensive
review on PEM water electrolysis. International Journal of Hydrogen
Energy, 38(12), pp.4901–4934.
doi:https://doi.org/10.1016/j.ijhydene.2013.01.151.

Cetinkaya, E., Dincer, I. and Naterer, G.F. (2012). Life cycle assessment of
various hydrogen production methods. International Journal of Hydrogen
Energy, 37(3), pp.2071–2080.
doi:https://doi.org/10.1016/j.ijhydene.2011.10.064.

Collodi, G., Azzaro, G., Ferrari, N. and Santos, S. (2017). Techno-economic


Evaluation of Deploying CCS in SMR Based Merchant H2 Production with
NG as Feedstock and Fuel. Energy Procedia, 114, pp.2690–2712.
doi:https://doi.org/10.1016/j.egypro.2017.03.1533.

El-Emam, R.S. and Özcan, H. (2019). Comprehensive review on the


techno-economics of sustainable large-scale clean hydrogen production.
Journal of Cleaner Production, 220, pp.593–609.
doi:https://doi.org/10.1016/j.jclepro.2019.01.309.
Energy.gov. Hydrogen Shot. [online] Available at:
https://www.energy.gov/eere/fuelcells/hydrogen-shot.

79
GOV.UK. Hydrogen production costs 2021. [online] Available at:
https://www.gov.uk/government/publications/hydrogen-production-costs-
2021

Hermesmann, M. and Müller, T.E. (2022). Green, Turquoise, Blue, or Grey?


Environmentally friendly Hydrogen Production in Transforming Energy
Systems. Progress in Energy and Combustion Science, 90, p.100996.
doi:https://doi.org/10.1016/j.pecs.2022.100996.

Hren, R., Vujanović, A., Van Fan, Y., Klemeš, J.J., Krajnc, D. and Čuček, L.
(2023). Hydrogen production, storage and transport for renewable energy
and chemicals: An environmental footprint assessment. Renewable and
Sustainable Energy Reviews, 173, p.113113.
doi:https://doi.org/10.1016/j.rser.2022.113113.

International Energy Agency (2021). Global Hydrogen Review 2021.

Ji, M. and Wang, J. (2021). Review and comparison of various hydrogen


production methods based on costs and life cycle impact assessment
indicators. International Journal of Hydrogen Energy, 46(78), pp.38612–
38635. doi:https://doi.org/10.1016/j.ijhydene.2021.09.142.

Kaplan, R. and Kopacz, M. (2020). Economic Conditions for Developing


Hydrogen Production Based on Coal Gasification with Carbon Capture and
Storage in Poland. Energies, 13(19), p.5074.
doi:https://doi.org/10.3390/en13195074.

Karaca, A.E., Dincer, I. and Gu, J. (2020). Life cycle assessment study on
nuclear based sustainable hydrogen production options. International
Journal of Hydrogen Energy, 45(41), pp.22148–22159.
doi:https://doi.org/10.1016/j.ijhydene.2020.06.030.
Katebah, M., Al-Rawashdeh, M. and Linke, P. (2022). Analysis of hydrogen
production costs in Steam-Methane Reforming considering integration with
electrolysis and CO2 capture. Cleaner Engineering and Technology, [online]
10, p.100552. doi:https://doi.org/10.1016/j.clet.2022.100552.

Khojasteh Salkuyeh, Y., Saville, B.A. and MacLean, H.L. (2017). Techno-
economic analysis and life cycle assessment of hydrogen production from

80
natural gas using current and emerging technologies. International Journal
of Hydrogen Energy, 42(30), pp.18894–18909.
doi:https://doi.org/10.1016/j.ijhydene.2017.05.219.

Kopacz, M., Kapłan, R. and Kwaśniewski, K. (2019). Evaluation of the


economic efficiency of hydrogen production by lignite gasification. Polityka
Energetyczna – Energy Policy Journal, 22(4), pp.21–36.
https://doi.org/10.33223/epj/113369

Kuckshinrichs, W., Ketelaer, T. and Koj, J.C. (2017). Economic Analysis of


Improved Alkaline Water Electrolysis. Frontiers in Energy Research, [online]
5. doi:https://doi.org/10.3389/fenrg.2017.00001.

Lee, S., Kim, H.S., Park, J., Kang, B.M., Cho, C.-H., Lim, H. and Won, W.
(2021). Scenario-Based Techno-Economic Analysis of Steam Methane
Reforming Process for Hydrogen Production. Applied Sciences, [online]
11(13), p.6021. doi:https://doi.org/10.3390/app11136021.
Li, J. and Cheng, W. (2020). Comparative life cycle energy consumption,
carbon emissions and economic costs of hydrogen production from coke
oven gas and coal gasification. International Journal of Hydrogen Energy,
45(51), pp.27979–27993.
doi:https://doi.org/10.1016/j.ijhydene.2020.07.079.

Li, J., Wei, Y.-M., Liu, L., Li, X. and Yan, R. (2022). The carbon footprint and
cost of coal-based hydrogen production with and without carbon capture
and storage technology in China. Journal of Cleaner Production, 362,
p.132514. doi:https://doi.org/10.1016/j.jclepro.2022.132514.
Li, Q., Song, G., Xiao, J., Hao, J., Li, H. and Yuan, Y. (2020). Exergetic life
cycle as-sessment of hydrogen production from biomass staged-
gasification. Energy, 190, p.116416.
doi:https://doi.org/10.1016/j.energy.2019.116416.
Lv, P., Wu, C., Ma, L. and Yuan, Z. (2008). A study on the economic
efficiency of hydrogen production from biomass residues in China.
Renewable Energy, 33(8), pp.1874–1879.
doi:https://doi.org/10.1016/j.renene.2007.11.002.

81
Mah, A.X.Y., Ho, W.S., Bong, C.P.C., Hassim, M.H., Liew, P.Y., Asli, U.A.,
Kamaruddin, M.J. and Chemmangattuvalappil, N.G. (2019). Review of
hydrogen economy in Malaysia and its way forward. International Journal of
Hydrogen Energy, [online] 44(12), pp.5661–5675.
doi:https://doi.org/10.1016/j.ijhydene.2019.01.077.
Massarweh, O., Al-khuzaei, M., Al-Shafi, M., Bicer, Y. and Abushaikha, A.S.
(2023). Blue hydrogen production from natural gas reservoirs: A review of
application and feasibility. Journal of CO2 Utilization, [online] 70, p.102438.

doi:https://doi.org/10.1016/j.ijhydene.2019.05.092.

Mehrpooya, M. and Habibi, R. (2020). A review on hydrogen production


thermochemical water-splitting cycles. Journal of Cleaner Production, 275,
p.123836. doi:https://doi.org/10.1016/j.jclepro.2020.123836.

doi:https://doi.org/10.1016/j.enconman.2010.10.023.
Muellerlanger, F., Tzimas, E., Kaltschmitt, M. and Peteves, S. (2007).
Techno-economic assessment of hydrogen production processes for the
hydrogen economy for the short and medium term. International Journal of
Hydrogen Energy, 32(16), pp.3797–3810.
https://doi.org/10.1016/j.ijhydene.2007.05.027

Netl.doe.gov. 7.1. Why Hydrogen? [online] Available at:


https://www.netl.doe.gov/research/Coal/energy-
systems/gasification/gasifipedia/hydrogen [Accessed 7 Jun. 2023].

Ni, M., Leung, D.Y.C., Leung, M.K.H. and Sumathy, K. (2006). An overview
of hydrogen production from biomass. Fuel Processing Technology, [online]
87(5), pp.461–472. doi:https://doi.org/10.1016/j.fuproc.2005.11.003.

Nikolaidis, P. and Poullikkas, A. (2017). A comparative overview of


hydrogen production processes. Renewable and Sustainable Energy
Reviews, [online] 67, pp.597–611.
doi:https://doi.org/10.1016/j.rser.2016.09.044.

Olateju, B., Kumar, A. and Secanell, M. (2016). A techno-economic


assessment of large scale wind-hydrogen production with energy storage in

82
Western Canada. International Journal of Hydrogen Energy, 41(21),
pp.8755–8776. doi:https://doi.org/10.1016/j.ijhydene.2016.03.177.
Orhan, M.F., Dincer, I. and Rosen, M.A. (2010). An exergy–cost–energy–
mass analysis of a hybrid copper–chlorine thermochemical cycle for
hydrogen production. International Journal of Hydrogen Energy, 35(10),
pp.4831–4838. doi:https://doi.org/10.1016/j.ijhydene.2009.08.095.

Parkinson, B., Balcombe, P., Speirs, J.F., Hawkes, A.D. and Hellgardt, K.
(2019). Levelized cost of CO2 mitigation from hydrogen production routes.
Energy & Environmental Science, 12(1), pp.19–40.
doi:https://doi.org/10.1039/c8ee02079e.

Poluzzi, A., Guandalini, G. and Romano, M.C. (2022). Flexible methanol


and hydrogen production from biomass gasification with negative
emissions. Sustainable Energy & Fuels.
doi:https://doi.org/10.1039/d2se00661h.

Ramsden, T., Steward, D. and J. Zuboy (2009). Analyzing the Levelized


Cost of Centralized and Distributed Hydrogen Production Using the H2A
Prodution Model, Version 2. doi:https://doi.org/10.2172/965528.

Salkuyeh, Y.K., Saville, B.A. and MacLean, H.L. (2018). Techno-economic


analysis and life cycle assessment of hydrogen production from different
biomass gasification processes. International Journal of Hydrogen Energy,
43(20), pp.9514–9528. doi:https://doi.org/10.1016/j.ijhydene.2018.04.024.
Shashank Reddy Patlolla, Katsu, K., Amir Sharafian, Wei, K., Herrera, O.E.
and Mérida, W. (2023). A review of methane pyrolysis technologies for
hydrogen production. 181, pp.113323–113323.

Spath, P.L. and Mann, M.K. (2000). Life Cycle Assessment of Hydrogen
Production via Natural Gas Steam Reforming. [online]
doi:https://doi.org/10.2172/764485.

Stiegel, G. J., & Ramezan, M. (2006, January). Hydrogen from coal


gasification: An economical pathway to a sustainable energy future.
International Journal of Coal Geology, 65(3–4), 173–190.
https://doi.org/10.1016/j.coal.2005.05.002

83
The Future of Hydrogen. (2019, June). IEA.

Utgikar, V. and Thiesen, T. (2006). Life cycle assessment of high


temperature electrolysis for hydrogen production via nuclear energy.
International Journal of Hydrogen Energy, 31(7), pp.939–944.
doi:https://doi.org/10.1016/j.ijhydene.2005.07.001.

Varbanov, P., Wang, Q., Zeng, M., Seferlis, P., Ma, T., Klemeš, J.,
Roussanaly, S., Anantharaman, R. and Fu, C. (2020). CHEMICAL
ENGINEERING TRANSACTIONS Low-Carbon Footprint Hydrogen
Production from Natural Gas: a Techno-Economic Analysis of Carbon
Capture and Storage from Steam-Methane Reforming. [online]
doi:https://doi.org/10.3303/CET2081170.

Yukesh Kannah, R., Kavitha, S., Preethi, Parthiba Karthikeyan, O., Kumar,
G., Dai-Viet, N.Vo. and Rajesh Banu, J. (2021). Techno-economic
assessment of various hydrogen production methods – A review.
Bioresource Technology, 319, p.124175.

El-Emam, R.S. and Khamis, I. (2017). International collaboration in the


IAEA nuclear hydrogen production program for benchmarking of HEEP.
International Journal of Hydrogen Energy, 42(6), pp.3566–3571.
doi:https://doi.org/10.1016/j.ijhydene.2016.07.256.

84

You might also like