Download as pdf or txt
Download as pdf or txt
You are on page 1of 114

Microwave-Assisted Methane Pyrolysis in a Fluidized Bed

Reactor for CO2-free Hydrogen and Solid Carbon Production

By

Mohammad Fawaz Khan

A thesis submitted in conformity with the requirements


for the degree of Master of Applied Science

Department of Mechanical & Industrial Engineering


University of Toronto

© Copyright by Mohammad Fawaz Khan 2023


Microwave-Assisted Methane Pyrolysis in a Fluidized Bed Reactor for CO2-free Hydrogen and Solid
Carbon Production

Mohammad Fawaz Khan

Master of Applied Science

Department of Mechanical & Industrial Engineering


University of Toronto

2023

Abstract
Hydrogen has the potential to decarbonize energy systems if it can become simpler to distribute

and cleaner to produce. Methane pyrolysis can produce hydrogen without direct CO2 emissions with

modest energy inputs. Microwaves are an efficient method of adding the required energy for this

endothermic reaction. This thesis investigates a novel method of low-GHG hydrogen production via

methane pyrolysis using microwaves and carbon particles in a fluidized bed reactor. Activated carbon

particles in the fluidized bed absorb microwave energy and create a hot medium (>1200℃) in contact

with flowing methane. As a result, methane decomposes into hydrogen and solid carbon causing carbon

catalyst deactivation with over 90% methane conversion rate and hydrogen selectivity. The produced

carbon possesses an autocatalytic effect thus bypassing the need for catalyst regeneration or replacement.

This modular pyrolysis system can be built anywhere with access to natural gas and electricity, enabling

distributed and localized hydrogen production.

ii
Acknowledgements
I would like to express my sincere gratitude and thanks to Professor Murray Thomson for taking me on
and providing me with this very interesting opportunity. His knowledge and guidance on the subject were
very instrumental in the success of this project.

I would also like to thank my fellow colleagues in the Thomson Lab, Mehran Dadsetan and Mehdi
Salakhi for their support and help. Mehran played an important role in the experimental setup and its
results. His mentorship allowed me to become a better researcher.

Thank you to Professor David Sinton for agreeing to be on my defense committee.

Lastly, I would like to thank Wayne Love for his technical expertise on microwaves and assisting us on
any problems that came up during this project.

iii
Statement of Contribution
This project was done in a group with other students so a statement of contribution on the work is made
below:

• Any modelling work presented is done by Mehdi Salakhi


• Any SEM images presented were taken by Mehran Dadsetan
• Experiments and corresponding results were carried out by Mohammad Fawaz Khan and Mehran
Dadsetan
• Chapter 4 is an edited version of a paper that is being submitted for publication

iv
Table of Contents
Abstract ......................................................................................................................................................... ii
Acknowledgements ...................................................................................................................................... iii
Statement of Contribution ............................................................................................................................ iv
Table of Contents .......................................................................................................................................... v
List of Tables ............................................................................................................................................. viii
List of Figures .............................................................................................................................................. ix
Chapter 1: Introduction ................................................................................................................................. 1
1.1 The Hydrogen Economy ..................................................................................................................... 1
1.2 Methane Pyrolysis Experimental Design Introduction ....................................................................... 2
1.3 Research Objectives ............................................................................................................................ 3
1.3 Thesis Outline ..................................................................................................................................... 4
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis ............ 5
2.1 Properties of Hydrogen ....................................................................................................................... 5
2.2 Applications of Hydrogen ................................................................................................................... 6
2.2.1 Oil Refining ................................................................................................................................. 8
2.2.2 Chemical Production .................................................................................................................... 9
2.2.3 Steel Production ......................................................................................................................... 10
2.2.4 Transportation Sector ................................................................................................................. 11
2.3 Hydrogen Production ........................................................................................................................ 12
2.3.1 Types of Hydrogen Production .................................................................................................. 13
2.3.2 Steam Methane Reforming ........................................................................................................ 15
2.3.3 Steam Methane Reforming + Carbon Capture and Sequestration or Carbon Capture and
Utilization ........................................................................................................................................... 16
2.3.4 Electrolysis ................................................................................................................................. 19
2.3.5 Methane Pyrolysis ...................................................................................................................... 21
2.3.5.1 Thermal Decomposition...................................................................................................... 25
2.3.5.2 Plasma Decomposition........................................................................................................ 27
2.3.5.3 Catalytic Decomposition ..................................................................................................... 28
2.3.5.3.1 Metal Catalysts............................................................................................................. 29
2.3.5.3.2 Carbon Catalysts .......................................................................................................... 29
2.4 Microwave Radiation ........................................................................................................................ 31
2.5 Fluidization ....................................................................................................................................... 33

v
Chapter 3: Experimental Apparatus ............................................................................................................ 37
3.1 Microwave Unit Preparation ............................................................................................................. 37
3.1.1 Microwave Unit ......................................................................................................................... 40
3.1.2 Microwave Chokes .................................................................................................................... 41
3.1.3 Reactor Vessel ........................................................................................................................... 42
3.1.4 Fritted Disc................................................................................................................................. 43
3.1.5 Flow Meters ............................................................................................................................... 44
3.1.6 Insulation.................................................................................................................................... 45
3.1.7 Sealing........................................................................................................................................ 46
3.1.8 Temperature Sensor ................................................................................................................... 46
3.1.9 Ventilation.................................................................................................................................. 47
3.1.10 Safety ....................................................................................................................................... 47
3.2 Fluidization ....................................................................................................................................... 48
3.2.1 Particle Size ............................................................................................................................... 48
3.2.2 Fluidization Regimes ................................................................................................................. 48
3.3 Carbon Particles ................................................................................................................................ 49
3.3.1 Carbon Particle Properties ......................................................................................................... 49
3.3.2 Grinding Particles ...................................................................................................................... 49
3.3.3 Sieving Particles......................................................................................................................... 49
3.4 Microwave Unit Preparation ............................................................................................................. 50
3.4.1 Cleaning the Microwave Cavity ................................................................................................ 50
3.4.2 Preparing and Installing Reaction Vessel .................................................................................. 50
3.4.3 Installing and Sealing Plumbing ................................................................................................ 51
3.4.4 Flow Controller Calibration ....................................................................................................... 51
3.4.5 Filtration ..................................................................................................................................... 52
3.5 Exhaust Analysis............................................................................................................................... 53
3.5.1 GC-TCD ..................................................................................................................................... 53
3.5.2 Total Chrom Navigator .............................................................................................................. 54
Chapter 4: Experimental Results................................................................................................................. 55
4.1 Carbon Particle Heating, Deactivation and Growth .......................................................................... 57
4.2 Methane Pyrolysis Conversion Rates ............................................................................................... 62
4.3 Major and Minor Species .................................................................................................................. 64
4.4 Carbon Mass Balance ....................................................................................................................... 65

vi
4.5 Thin-Film Carbon ............................................................................................................................. 66
4.6 Methane Dilution .............................................................................................................................. 68
4.7 Activation Energy ............................................................................................................................. 69
Chapter 5: Conclusion and Future Work .................................................................................................... 70
5.1 Conclusion ........................................................................................................................................ 70
5.2 Future Work ...................................................................................................................................... 71
5.2.1 Experimental Apparatus Improvements ..................................................................................... 71
5.2.2 Future Research.......................................................................................................................... 71
References ................................................................................................................................................... 73
Appendix A Electrolysis Technologies....................................................................................................... 85
A.1 Alkaline Water Electrolysis (AWE)................................................................................................. 85
A.2 Proton Exchange Membrane Electrolysis (PEM) ............................................................................ 86
A.3 Solid Oxide Electrolysis (SOEC) ..................................................................................................... 87
Appendix B Experiment Setup Procedure .................................................................................................. 88
Appendix C Experiment Operation Procedure ........................................................................................... 90
Appendix D GC-TCD Operation Procedure ............................................................................................... 91
Appendix E Fused Quartz Tube Properties................................................................................................. 97
Appendix F Flow Controller Calibration Procedure ................................................................................... 98
Appendix G SEM Images of Carbon Granules ........................................................................................... 99
Appendix H Inception Model, Simulation Conditions and Surface Growth ............................................ 102
Appendix I Effective Activation Energy Calculation ............................................................................... 103

vii
List of Tables
Table 1. Properties of hydrogen with comparison to other fuels .................................................................. 6
Table 2. Hydrogen production technologies with primary feedstock, technology readiness level (TRL) and
level of CO2 emissions ............................................................................................................................... 14
Table 3. Energy required to produce hydrogen in SMR, electrolysis and methane pyrolysis .................... 22
Table 4. Carbon mass balance experiments under two different inlet mixture and temperature conditions
.................................................................................................................................................................... 65
Table 5. Quartz tube properties from Technical Glass Products, Inc. ........................................................ 97
Table 6. Table displaying values inputted in Brooks mass flow controller with resulting flow rate shown
in Bios. From this, the calibration curve in equation (25) was determined. ............................................... 98
Table 7. Simulation conditions. The geometry, boundary conditions and sectional model parameters used
in the NanoPFR code are shown here ....................................................................................................... 102

viii
List of Figures
Figure 1. Hydrogen production by microwave heating of a fluidized activated carbon particle bed ........... 3
Figure 2. Global demand for pure hydrogen in Mt, 1975-2018 .................................................................... 7
Figure 3. Global hydrogen demand in Mt by sector in the Net Zero Scenario, 2020-2030 .......................... 7
Figure 4. Hydrogen applications ................................................................................................................... 8
Figure 5. Allowed sulfur content in refined products versus oil demand ..................................................... 9
Figure 6. Hydrogen demand for primary chemical production for existing applications under current
trends. MTO = methanol-to-olefins; MTA = methanol-to-aromatics ......................................................... 10
Figure 7. Hydrogen production technologies .............................................................................................. 13
Figure 8. Steam-methane reforming process .............................................................................................. 16
Figure 9. Carbon capture and storage or utilization options ....................................................................... 17
Figure 10. Carbon capture options in SMR process ................................................................................... 18
Figure 11. Principles of an electrolyzer ...................................................................................................... 20
Figure 12. Reactor types with heat source and operating temperature range ............................................. 23
Figure 13. Types of carbon produced expressed as a function of types of catalyst and temperature ......... 24
Figure 14. Categories of methane pyrolysis................................................................................................ 25
Figure 15. Schematic of methane pyrolysis in a liquid metal bath ............................................................. 26
Figure 16. Schematic of methane pyrolysis in a fluidized bed reactor composed of carbon granules ....... 27
Figure 17. Schematic of plasma pyrolysis reactor ...................................................................................... 28
Figure 18. Electromagnetic spectrum ......................................................................................................... 32
Figure 19. Fluidization regimes .................................................................................................................. 34
Figure 20. Geldart's classification of fluidized particles ............................................................................. 35
Figure 21. Schematic of experimental setup for microwave-driven methane pyrolysis: (A) Methane and
nitrogen cylinders, (B) microwave chokes, (C) fritted discs, (D) 1000 W microwave, (E) thermocouple,
(F) muffle, (G) reactor filled with carbon particles, and (H) GC – TCD .................................................... 38
Figure 22. Experimental setup for microwave-driven methane pyrolysis: (A) Vent, (B) H2 sensor, (C)
thermocouple, (D) exhaust line, (E) microwave, (F) inlet line, (G) pressure gauge, (H) pressure relief
valve ............................................................................................................................................................ 39
Figure 23. 1 kW, 2.450 GHz microwave unit ............................................................................................. 40
Figure 24. Microwave chokes attached to microwave (left) and detached (right) ...................................... 41
Figure 25. Quartz reaction vessel ................................................................................................................ 42
Figure 26. Fritted discs 20 mm in diameter ................................................................................................ 43

ix
Figure 27. Muffle insulation with no door (left) and door (right) ............................................................... 45
Figure 28. Ultra torr sealing Swagelok connection ..................................................................................... 46
Figure 29. Ventilation fan ........................................................................................................................... 47
Figure 30. Quartz vessel filled with carbon particles sitting on fritted discs .............................................. 50
Figure 31. Ultra torr sealing configuration for bottom (left) and top (right) choke .................................... 51
Figure 32. Filter paper (left) and 2-micron particulate filter (right) are used to filter the exhaust.............. 52
Figure 33. GC-TCD .................................................................................................................................... 53
Figure 34. Exhaust results on Total Chrom Navigator showing hydrogen, nitrogen and methane peaks .. 54
Figure 35. Hydrogen production by microwave heating of a fluidized activated carbon particle bed. (A)
Vertical microwave with fluidized bed inside. (B) Fluidized bed reactor where methane flows from
bottom to top. (C) Carbon particles absorb microwave radiation and heat up for the methane pyrolysis
reaction. At temperatures above 950℃, methane starts decomposing to hydrogen and solid carbon. The
initial particles grow in size due to carbon deposition onto their surface. .................................................. 56
Figure 36. Effect of carbon deposition on the surface of activated carbon expressed by time versus
methane conversion rate. Methane conversion rate starts at ~100%, but quickly declines reaching a steady
state after almost three hours of run time. ................................................................................................... 58
Figure 37. Scanning electron microscope (SEM) images of activated carbon particles (212 - 425 microns)
before pyrolysis (A) and after pyrolysis (B) from 1mm, 300 micrometer and 50 micrometer magnification
respectively ................................................................................................................................................. 59
Figure 38. Carbon particle agglomerates from carbon deposition bridging particles due to poor
fluidization .................................................................................................................................................. 60
Figure 39. Gas temperature in contact with covered particles absorbing microwave energy ..................... 60
Figure 40. Heated quartz reactor ................................................................................................................. 61
Figure 41. Methane conversion rates and hydrogen selectivity over a range of temperatures, pressure of 1
atm and estimated residence time of ~5 seconds ........................................................................................ 63
Figure 42. Major species (left) and minor species (right) from methane pyrolysis over a range of
temperatures. The symbols represent the experimental results and the solid, dash and dotted lines
represent the modeling results .................................................................................................................... 64
Figure 43. Thin film carbon formation in reactor ....................................................................................... 66
Figure 44. Collected thin film carbon ......................................................................................................... 67
Figure 45. SEM image of thin film carbon at 50 micrometers ................................................................... 67
Figure 46. Effect of dilution on methane conversion rate at a pressure of 1 atm and estimated residence
time of ~5 seconds ...................................................................................................................................... 68

x
Figure 47. Activation energy of covered carbon particles at temperatures of 600 – 900°C, pressure of 1
atm and estimated residence time of ~5 seconds ........................................................................................ 69

Figure A1. Operating principle of alkaline water electrolysis .................................................................... 85


Figure A2. Operating principle of proton exchange membrane electrolysis cell ........................................ 86
Figure A3. Operating principle of solid oxide electrolysis cell .................................................................. 87

Figure B1 .................................................................................................................................................... 88
Figure B2 .................................................................................................................................................... 88
Figure B3 .................................................................................................................................................... 89
Figure B4 .................................................................................................................................................... 89

Figure D1 .................................................................................................................................................... 91
Figure D2 .................................................................................................................................................... 91
Figure D3 .................................................................................................................................................... 92
Figure D4 .................................................................................................................................................... 93
Figure D5 .................................................................................................................................................... 94
Figure D6 .................................................................................................................................................... 95

Figure G1. Activated carbon before methane pyrolysis at 50 micrometer scale ........................................ 99
Figure G2. Activated carbon before methane pyrolysis at 30 micrometer scale ........................................ 99
Figure G3. Activated carbon after methane pyrolysis at 50 micrometer scale ......................................... 100
Figure G4. Activated carbon after methane pyrolysis at 30 micrometer scale ......................................... 100
Figure G5. Thin film carbon at 300 micrometer scale .............................................................................. 101
Figure G6. Thin film carbon at 50 micrometer scale ................................................................................ 101

xi
Chapter 1: Introduction

Chapter 1: Introduction
1.1 The Hydrogen Economy
Hydrogen appears to be gaining a significant amount of attention over the past few decades by the
media and scientists alike. It’s been dubbed the “fuel of the future” [1] with the conversation not too far
from the term, “the hydrogen economy” [2]. With a lot of buzz around hydrogen, it’s important to ask
questions like ‘what is the hydrogen economy? why is it important?’. The hydrogen economy is described
as a utopic vision of society where the energy and economic infrastructure is dependent on hydrogen fuel
rather than the fossil fuels used today: oil, coal, and natural gas. The importance is its potential ability to
transition away from the reliance on fossil fuels and mitigate climate change. This stems from the reality
that hydrogen is a clean fuel when combusted and is widely available in other hydrogen carrying forms
such as water or hydrocarbons. The IPCC have stated that to limit global warming by 2°C, greenhouse
gas (GHG) emissions should be reduced by 25% by around 2030 [3].

Hydrogen is not found naturally, therefore it must be produced through other means. Currently,
it’s produced from steam methane reforming (SMR) alongside CO2 as a by-product. Numerous technical
challenges hold hydrogen back from becoming the dominant form of fuel in the global economy. One of
the issues being its method of production. A clean method of production is needed in order for hydrogen
to be truly considered a clean fuel. Other issues surrounding hydrogen are, but not limited to, economic
competitiveness of current clean methods to produce hydrogen, difficulty in development of storage and
transportation of hydrogen, flammable safety concerns and more.

If a hydrogen economy were to ever be realized, a clean method to produce it must be


economically competitive to SMR. Methods to produce clean hydrogen have been proposed such as
electrolysis where water is split into its fundamental components, oxygen and hydrogen. Another is
carbon capture and sequestration (CCS) where an additional process is applied near the end of SMR
where the CO2 is captured and sequestered underground. An alternative method for large-scale clean

1
Chapter 1: Introduction

hydrogen production has been of growing interest. This process is known as methane pyrolysis, the
decomposition of methane under extreme heat into its fundamental components, hydrogen and solid
carbon. This type of hydrogen production has been termed as turquoise hydrogen. The advantages of
methane pyrolysis are that its significantly less energy intensive compared to electrolysis, produces a
marketable carbon product and benefits from pre-existing natural gas infrastructure. Methane pyrolysis, is
believed to be superior to electrolysis and a potential replacement in fossil fuel-based hydrogen
production in order to achieve a hydrogen based economy [4]–[7]. The motive for this thesis is the design
and testing of a methane pyrolysis experimental apparatus to continuously produce CO2-free turquoise
hydrogen.

1.2 Methane Pyrolysis Experimental Design Introduction


Methane pyrolysis involves the thermal decomposition of methane molecules at elevated
temperatures of about 1000°C [5]. In the proposed design, methane pyrolysis will be driven by
microwave heating of activated carbon particles fluidized inside a quartz reactor (see Figure 1). The
methane will flow in from the bottom and into the quartz reaction vessel inside the microwave (Figure
1B). The carbon particles will be heated from the absorption of microwave energy to temperatures higher
than 1000°C causing the methane to decompose into hydrogen and solid carbon (Figure 1C). Microwaves
are a contactless and efficient form of heating compared to conventional heat transfer [8] so that the heat
can be generated directly where required [9]. A fluidized bed is implemented for its efficient particle to
gas heat transfer capabilities [10]. This pyrolysis design is quite modular and can be deployed anywhere
as long as there is a natural gas and electricity source readily available. The modularity helps provide
localized hydrogen production ultimately avoiding the hydrogen distribution and storage challenges that
are commonly associated with the handling of hydrogen [11].

2
Chapter 1: Introduction

Figure 1. Hydrogen production by microwave heating of a fluidized activated carbon particle bed

1.3 Research Objectives


The ultimate objective of this research is to produce CO2-free hydrogen from the pyrolysis of
methane. This will be accomplished with the design and construction of a lab-scale microwave methane
pyrolysis unit in a fluidized bed reactor. Upon the completion of the experimental setup, the following
experimental objectives are sought to be achieved:

(A) Methane conversion rate


(B) Hydrogen selectivity
(C) Major and minor species that appear in methane decomposition
(D) Effect of methane dilution
(E) Carbon catalyst deactivation
(F) Effect of fluidization on methane conversion rate
(G) Carbon particle growth and morphology study before and after methane pyrolysis
(H) Limitations of the experimental setup with suggestions on improvements

3
Chapter 1: Introduction

1.3 Thesis Outline


In this thesis, chapter 2 discusses background information that may be of interest with regards to
the thermodynamics and basic principles associated with the design proposed in the section above. Some
topics that are discussed in chapter 2 are current applications of hydrogen, hydrogen production methods,
microwave energy and fluidization. Chapter 3 discusses the experimental apparatus to produce hydrogen
via microwave driven methane pyrolysis. Chapter 4 discusses the results of the experiments. Chapter 5
concludes this thesis along with a discussion on future work.

4
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Chapter 2: Hydrogen in the Energy


Landscape and the Influencing Factors of
Methane Pyrolysis
2.1 Properties of Hydrogen
No paper discussing hydrogen can be written without stating that it’s the most abundant and
simplest element in the universe consisting of one proton and electron. Hydrogen does not freely exist in
nature and must be extracted from the many different forms it’s stored in such as in water (H2O) and other
hydrocarbons such as methane (CH4). Hydrogen is a combustible fuel that gives off heat and water vapor
when combined with oxygen as seen in equation (1). It does not produce carbon dioxide unlike its
conventional fossil-fuel based counterparts.

1
𝐻2(𝑔) + 𝑂2(𝑔) → 𝐻2 𝑂(𝑙) (1)
2

The heat of combustion of hydrogen at 25°C and 1 atm is equal to the heat of formation of water,
286 kJ/mol (142 MJ/kg) [12]. Since the hydrogen molecule is very small and possesses a low volumetric
energy density, there are challenges associated with storing and transporting the element. To store
hydrogen, it must be pressurized in its gaseous state or liquified at -252.76°C [12]. This is very costly to
establish on a large-scale level and challenges like these are necessary to overcome to establish a
hydrogen economy.

Despite the attractiveness of using hydrogen as a potentially clean fuel, it has a host of negative
aspects associated with it such as storage, transportation, low volumetric energy density and expensive to
produce from clean sources to name a few. The high cost in producing hydrogen from renewable sources
dissuades users from consuming it as since it’s far easier and cheaper to directly burn fossil fuels instead.

5
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Table 1 below shows the properties of hydrogen compared to fossil fuels such as natural gas, liquified
natural gas (LNG), gasoline and methane [13].

Table 1. Properties of hydrogen with comparison to other fuels [13]

Property of hydrogen Value Comparison with other fuels

Density (gas) 0.089 kg/m3 (0°C, 1 bar) 1/10 of natural gas

Density (liquid) 70.79 kg/m3 (-253°C, 1 bar) 1/6 of natural gas

Boiling point -252.76°C (1 bar) 90°C below LNG

HHV 142 MJ/kg [12] 3x that of gasoline

LHV 120.1 MJ/kg 3x that of gasoline

Energy density (ambient 0.01 MJ/L 1/3 of natural gas


conditions, LHV)

Liquified specific energy (LHV) 8.5 MJ/L 1/3 of LNG

Flame velocity 346 cm/s 8x methane

Ignition range 4 – 77 vol% in air 6x wider than methane

Autoignition temperature 585°C 220°C for gasoline

Ignition energy 0.02 MJ 1-10 of methane

Cost of production* $7.04/GJ – $12.32/GJ [14] $2.84/GJ – $10.43/GJ [14]

* - Cost of hydrogen is dependant on cost of natural gas, which varies greatly by region.

2.2 Applications of Hydrogen


Although hydrogen is not mainly used as a fuel in today’s world, it does have its applications and
boasts a very large market. The global demand for hydrogen has increased quite considerably over the
years. The hydrogen demand has shot up from just under 20 Mt in 1975 to just over 70 Mt in 2018 [14].
From Figure 2, it can be seen that this is a continuous increase in demand and consumption of hydrogen
show no signs of stopping. By 2030, the demand is expected to increase even more requiring about 200
Mt of hydrogen [15].

6
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Figure 2. Global demand for pure hydrogen in Mt, 1975-2018 [14]

250

200
Million tonnes of hydrogen (Mt)

150

100

50

0
2020 2025 2030

Refining Industry Transport Power Ammonia - fuel Synfuels Buildings Grid injection

Figure 3. Global hydrogen demand in Mt by sector in the Net Zero Scenario, 2020-2030 [15]

Most hydrogen is used to refine oil, which consists of 33% of hydrogen consumption [14]. 27%
and 11% are allocated to the production of ammonia and methanol respectively [14]. 3% of it is used for
steel production via the direct reduction of iron ore [14]. As can be seen in Figure 3, it is estimated that
the demand for hydrogen will double by 2030, specifically in the cases for ammonia production and grid
injection [15]. Minor uses for hydrogen are in other metallurgical and chemical production applications

7
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

and also as a fuel such as a rocket propellant. Many of the applications that hydrogen is involved in can
be seen in Figure 4 below [13].

Figure 4. Hydrogen applications [13]

2.2.1 Oil Refining


The oil refining industry uses 38 MtH2/yr accounting for 33% of the global hydrogen demand
[14]. Hydrogen is used as an agent for hydrocracking and hydrotreatment processes. Hydrocracking
involves the process of breaking down heavier hydrocarbon fuels into lighter and higher value
components such as gasoline, kerosene and diesel. This involves high temperatures and pressures along
with hydrogen and a catalyst [16].

Hydrotreatment consists of removing impurities within crude oil such as sulfur to reduce SOx
emissions that contribute to poor air quality. The process of removing sulfur using hydrogen is also
known as desulfurization. Over time, regulations on the sulfur content within crude oil (and other fossil
fuels) is getting tighter with an ever-increasing push on decreasing harmful emissions into the
environment [14]. This increase in tightening sulfur regulations will consequentially require a higher
demand in hydrogen.

8
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

In Figure 5, the oil demand has constantly increased over time and will continue to increase in the
near future [14], [17]–[19]. As a result of an increase in oil demand, the demand for hydrogen would also
increase. Hydrogen in refineries contribute to about 20% of the world’s total CO2 emissions [14]. With
the growing concern over large-scale fossil fuel consumption and its impact on global warming, it is
predicted that oil demand will eventually plateau or decline [15], [17]–[19]. It is up to debate as to
if/when a decline will start due to the many political, economic and technological barriers in the way.
Regardless of how the demand for oil will settle in the future, the future demand for hydrogen use in the
oil refining industry is directly dependent on the demand for future oil and sulfur content regulations.

Figure 5. Allowed sulfur content in refined products versus oil demand [14]

2.2.2 Chemical Production


Hydrogen is essential in the production of chemicals such as ammonia (NH3) and methanol
(CH3OH). Ammonia is produced using the Haber-Bosch process where hydrogen is combined with
nitrogen to make ammonia as seen in equation (2) [20]. The production of methanol involves the
hydrogenation of CO and CO2 in addition to the reversed water gas-shift reaction as seen in equations (3),
(4) and (5) respectively [21]. The ammonia and methanol production industry uses about 31 MtH2/yr and
12 MtH2/yr respectively, making up 38% of total global hydrogen demand [14]. Ammonia and methanol
are primarily used to manufacture fertilizers such as urea (CO(NH2)2) or ammonium nitrate (NH4NO3).
Urea and ammonium nitrate are agents in plastics, paints, dyes, explosives and other chemicals. Methanol
is used in the production of various industrial chemicals such as formaldehyde (CH2O) and methyl
methacrylate (C5H8O2) which are involved in resins, adhesives, dyes, disinfectants, paints and other
applications.

9
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

𝑁2 + 3𝐻2 → 2𝑁𝐻3 , ∆𝐻° = −91.8 𝑘𝐽/𝑚𝑜𝑙 (2)

𝐶𝑂 + 2𝐻2 ↔ 𝐶𝐻3 𝑂𝐻, ∆𝐻° = −90.55 𝑘𝐽/𝑚𝑜𝑙 (3)


𝐶𝑂2 + 3𝐻2 ↔ 𝐶𝐻3 𝑂𝐻 + 𝐻2 𝑂, ∆𝐻° = −49.43 𝑘𝐽/𝑚𝑜𝑙 (4)
𝐶𝑂2 + 𝐻2 ↔ 𝐶𝑂 + 𝐻2 𝑂, ∆𝐻° = 41.12 𝑘𝐽/𝑚𝑜𝑙 (5)

The IEA predicts that the future demand for ammonia and methanol in current applications will
grow by 31% primarily caused by economic and population growth by 2030 [14]. Figure 6 shows that the
demand for ammonia has increased historically over time and will continue to increase along with the
human population rate to accommodate the larger population by producing more food.

Figure 6. Hydrogen demand for primary chemical production for existing applications under current trends. MTO = methanol-to-
olefins; MTA = methanol-to-aromatics [14]

The long-term demand for hydrogen-based chemicals is primarily dependant on the demand for
their current respective applications. In 2050, it may be possible to observe a new use case for ammonia
or other chemicals as a zero-carbon emitting fuel or an entirely new application that may affect future
demand.

2.2.3 Steel Production


Steel is an alloy composing of primarily iron and some carbon. Iron is found naturally in the earth
as iron ore and is obtained by mining. Most iron ore is hematite, an iron oxide, with the formula, Fe2O3.
To produce the iron, the iron ore is sent into a blast furnace-basic oxygen furnace (BF-BOF) process
where it produces pig iron (liquid iron saturated in carbon) that have been stripped of its oxygen atoms.
This process accounts for about 90% of global primary steel production and its chemical reaction can be
seen below in equation (6) [14], [22].

10
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

2𝐹𝑒2 𝑂3 + 3𝐶 → 2𝐹𝑒 + 3𝐶𝑂2 (6)

The other process (the one of more significance for this thesis) is the direct reduction of iron
(DRI) process. This is where a mixture of hydrogen and carbon monoxide is used as a working agent to
remove the oxygen in iron oxide to produce pure iron as is displayed in equation (7) and equation (8)
[22]. The iron is melted for use in an electric arc furnace (EAF) to produce steel [22], [23]. The DRI-EAF
process accounts for 7% of global primary steel production and consumes about 4 MtH2/yr which is 3%
of total global hydrogen demand [14].

𝐹𝑒2 𝑂3 + 3𝐶𝑂 → 2𝐹𝑒 + 3𝐶𝑂2 (7)

𝐹𝑒2 𝑂3 + 3𝐻2 → 2𝐹𝑒 + 3𝐻2 𝑂 (8)

Steel demand is expected to increase over time [14]. DRI is not the dominant method for
producing steel, but it is thought of as a potential method to help lower CO2 emissions by using almost
100% pure hydrogen instead of the carbon monoxide and hydrogen mixture [14], [22], [24]. The total
hydrogen demand in the steel industry would be dictated by how much of a market share DRI-EAF could
capture versus the conventional BF-BOF.

2.2.4 Transportation Sector


Hydrogen is speculated to be a potential fuel in applications in the transportation industry such as
cars, trucks, buses, rail and planes. This is a very niche area of application mainly in the R&D stage today
as there are some issues with each of these potential applications. The issues surrounding these
applications mainly surround the transportation and storage issues of hydrogen requiring high pressures
and a temperature of -253°C which can be very costly. Concerning the auto industry, cars based on
hydrogen fuel cells face a “chicken or the egg” problem. No automaker wants to build hydrogen fuel cell
cars if there are no hydrogen refueling stations and no governing body wants to build hydrogen-refueling
stations if there are not any hydrogen cars. This is unlike electric vehicles where every North American
home has access to electricity and the prospect of becoming its own recharging station. Today, hydrogen
does have a niche application as a fuel in rocket propulsion. Whether or not the transport sector transitions
to hydrogen-based fuels remains to be seen.

A lot of the future demand for hydrogen depends on the future demand for its downstream
products in oil refining, fertilizers, and steel to name a few [14]. The demand for many of the downstream
products of hydrogen consumption is expected to increase and therefore, increase the global demand for
hydrogen. New applications for downstream products of hydrogen and technological breakthroughs that
11
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

effect any of the discussed above is entirely possibly but is difficult to predict and take into consideration.
Although there is a predicted rise in hydrogen demand, hydrogen does have its host of challenges when
handling it. One of those challenges is becoming more contentious with the now increased concern for
climate change: its means of production.

2.3 Hydrogen Production


Hydrogen is not found naturally in the world. It must be made from a hydrogen carrying
feedstock such as natural gas, oil, water, biomass etc. There are multiple different technologies that
attempt to capitalize on the abundant sources for hydrogen production ranging from fossil fuels to
renewable energy as shown in Figure 7 [25]. Natural gas, oil and coal can all be used to produce hydrogen
with natural gas and coal being the primary method of production. Most hydrogen is produced from fossil
fuels such as natural gas (76%) and coal (33%) producing about 830 MtCO2/yr [14]. Renewable
technologies to produce hydrogen, such as electrolysis, are not widely used and have their own set of
issues. One of these issues is the large energy consumption in electrolysis with the additional requirement
for the electrical energy to be sourced from renewable sources to be considered clean hydrogen.
Electrolysis is not the dominant form of hydrogen production and only contributes to about 2% of the
global hydrogen demand [14]. There are multiple different types of hydrogen production that will be
described in the following section.

12
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Steam reforming

Natural gas Partial oxidation

Autothermal reforming

Fossil Pyrolysis
Oil
resources
Gasification

Combustion
Coal
Liquefaction

Dark fermentation

Photo-fermentation
Biomass

Renewable Bio-photolysis
resources
Electrolysis

Water Thermolysis

Photolysis
Figure 7. Hydrogen production technologies [25]

2.3.1 Types of Hydrogen Production


The means of hydrogen production has been associated with a colour to illustrate the cleanliness
of a method with regards to greenhouse gas emissions. Steam-methane reforming (SMR) is the most
common form of hydrogen production but also emits CO2 and is dubbed “gray” hydrogen [14].

SMR combined with carbon capture and sequestration (CCS) technology is where the CO2
emissions from SMR is captured and stored instead of being emitted into the atmosphere. This process
does add an additional layer of cost but results in lower CO2 emissions and is termed “blue” hydrogen.

“Green” hydrogen involves hydrogen production by using renewable resources and thus zero CO2
emissions. The most common method of green hydrogen production is electrolysis where the electricity
comes from renewable sources. Electrolysis involves the splitting of water atoms into its fundamental
components, hydrogen and oxygen, by exciting them with electrical energy.

13
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Electricity produced from nuclear power has been used in electrolysis as well. Since nuclear
power isn’t entirely considered renewable energy, electrolysis containing electricity sourced from nuclear
power is known as “pink” hydrogen.

A rising form of hydrogen production is methane pyrolysis. This is where methane (CH4) is split
into solid carbon and hydrogen gas by heating up methane molecules. This process can be zero-
greenhouse gas emitting if the method of heating is sourced from renewable energy and no methane leaks
in the process. It’s been classified as in between green and blue, and has hence been termed, “turquoise”
hydrogen.

Each hydrogen production technology has its advantages and disadvantages that make them
superior to one another in certain conditions. Table 2 below shows how gray, blue, turquoise and green
hydrogen compare against one another showcasing its feedstock, production technology, technology
readiness level (TRL) and process-related CO2 emissions [25]. SMR, SMR + CCS, and PEM electrolysis
all have high TRL levels, but the cost of hydrogen produced through SMR is far cheaper than the other
methods thus being the most dominant form of hydrogen production today [26]. Turquoise hydrogen is
currently in the research and development phase and has been given a TRL of 3-4. However, some
companies have plans for large-scale commercial plants for turquoise hydrogen production, the TRL level
can be updated to about 6 [27].

Table 2. Hydrogen production technologies with primary feedstock, technology readiness level (TRL) and level of CO 2
emissions [25]

Associated
“Gray” “Blue” “Turquoise” “Green”
Colour

Primary
Natural Gas Natural Gas Natural Gas Water
Feedstock
Polymer
SMR with Carbon
Electrolyte
Production Steam-Methane Capture &
Methane Pyrolysis Membrane Water
Technology Reforming (SMR) Sequestration
Electrolysis
(CCS)
(PEMEL)
Research &
Technology Commercial Industrial Commercial
Development
Readiness Level (TRL 9) (TRL 8-9) (TRL 9)
(TRL 3-4)

Process-Related
High-CO2 Low-CO2 CO2-free Carbon-free
CO2 Emissions

14
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

2.3.2 Steam Methane Reforming


Steam-methane reforming (SMR) is the most common method used today to produce hydrogen
[14]. SMR consists of the combination of methane and steam over a catalyst to create hydrogen (H2) and
carbon dioxide (CO2) as shown in equation (9). This equation is an endothermic reaction requiring
temperatures in the range of 700 – 1000°C and up to 3.5 MPa to maintain a high CH4 to H2 conversion
[13], [28], [29]. However, SMR is not as simple as equation (9) and is a multi-step process as shown in
equations (10) to (11). The methane in the steam reforming process first takes the form of carbon
monoxide (CO) and hydrogen as displayed in equation (10). Equation (10) is also more favorable to occur
at high temperatures and low pressures. The CO undergoes the water-gas shift reaction in equation (11) to
make CO2 and H2.

Global Reaction:

𝐶𝐻4 (𝑔) + 2𝐻2 𝑂(𝑙) ↔ 𝐶𝑂2 (𝑔) + 4𝐻2 (𝑔), ∆𝐻° = 253.03 𝑘𝐽/𝑚𝑜𝑙 (9)

Steam Reforming Reaction:

𝐶𝐻4 (𝑔) + 𝐻2 𝑂(𝑙) ↔ 𝐶𝑂(𝑔) + 3𝐻2 (𝑔), ∆𝐻° = 250.2 𝑘𝐽/𝑚𝑜𝑙 (10)

Water-Gas Shift Reaction:

𝐶𝑂(𝑔) + 𝐻2 𝑂(𝑙) ↔ 𝐶𝑂2 (𝑔) + 𝐻2 (𝑔), ∆𝐻° = 2.83 𝑘𝐽/𝑚𝑜𝑙 (11)

Other Intermediate Reactions, but not limited to:

𝐶𝐻4 → 𝐶 + 2𝐻2 , ∆𝐻° = 74.87 𝑘𝐽/𝑚𝑜𝑙 (12)


2𝐶𝑂 → 𝐶𝑂2 + 𝐶, ∆𝐻° = −172.5 𝑘𝐽/𝑚𝑜𝑙 (13)

Figure 8 shows the general process of SMR [25]. Natural gas is the primary feedstock and
consists mainly of methane ranging from a mole percentage of 87 – 98% with trace amounts of minor
species such as, but not limited to, ethane, propane, butane and sulfur in its composition [30]. Sulfur and
other contaminants can harm and deactivate the catalysts in the process [29], [31]. To protect the
catalysts, the incoming natural gas will go through a pre-treatment process to remove any contaminants
such as hydrogen sulfide (H2S) by hydrotreating using a zinc oxide (ZnO) catalyst at about 340 – 390°C
[32].

After pre-treatment of the natural gas, the steam reforming process begins by undergoing the
steam reforming reaction in equation (10) where natural gas is combined with high temperature steam

15
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

using a nickel catalyst [28], [29], [32]–[34]. A high operating temperature is required due to the
endothermic nature of the reaction [28]. From equation (12) and (13), coke from methane decomposition
and carbon monoxide can deposit on the catalyst and deactivate it. To avoid this, SMR operates with
excess steam to promote more output products (CO and H2) at a steam-to-methane molar ratio of 2.5 – 5
[28], [29], [32].

The CO from the steam reforming reaction will undergo the water-gas shift reaction to convert
the CO into CO2 and H2 as per equation (11) using a two-stage catalytic reactor. The hydrogen is then
purified by pressure swing absorption (PSA) where the hydrogen is separated from all the exhaust.

Figure 8. Steam-methane reforming process [25]

2.3.3 Steam Methane Reforming + Carbon Capture and Sequestration or


Carbon Capture and Utilization
With the increasing concern for carbon emissions released into the atmosphere, there has been an
ever-growing interest in trying to mitigate the carbon emissions from SMR. One way is to add an
intermediate step where the CO2 is captured and sequestered before it can be released into the
atmosphere, known as carbon capture and sequestration (CCS). On the other hand, instead of
sequestration, the CO2 can be utilized as different products through conversion, known as carbon capture
and utilization (CCU). If a large percentage of CO2 is captured and stored or utilized, this would become a
low-CO2 process and is termed “blue” hydrogen.

There are many different industries that release CO2 capable of applying CCS technology such as
the natural gas, coal and steel industry. Due to the diversity of CO2 emitting processes, different designs
to capture CO2 are appropriately required. Figure 9 below provides a summary of CCS and CCU options
and pathways to answer this problem [35]. For example, some CO2 removal technologies are, but not
limited to, physical or chemical absorption, adsorption, membrane, cryogenic separation or pressure
swing adsorption (PSA) [35], [36]. Figure 10 shows the conventional SMR process with the addition of

16
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

options where CO2 capture can take place [25], [26]. CO2 removal in a SMR plant can be captured in
three different ways and is described further below: pre-combustion removal (Option 1), tail gas removal
(Option 2) and post-combustion removal in the flue gas (Option 3) [25], [26], [36], [37].

1. Pre-combustion (or pre-conversion) CO2 removal refers to the separation and capture of CO2
contained within the feedstock or from an intermediate process. Considering the case of SMR,
after CO conversion into CO2 from the water-gas shift reaction, the CO2 is removed through
absorption by chemical or physical absorption by the industry standard methyldiethanolamine
(MDEA) [26], [37]. The CO2 is available at higher concentrations and pressures of 16%v CO2 at
2.5 MPa respectively and allows for simpler separation [26], [36].
2. CO2 separation from tail gas from PSA using MDEA or cryogenic and membrane separation [26].
This is more difficult to do because of its low pressure at 0.2 MPa [26].
3. Post-combustion CO2 removal is the removal of CO2 from flue gas line after the carbon source
has been converted to CO2 [36]. This occurs at low pressures of 0.1 MPa and involves PSA in
combination with an adsorbent such as monoethanolamine (MEA) or MDEA, cryogenic and/or
membrane separation to remove the CO2 at higher pressures [26], [35].

Figure 9. Carbon capture and storage or utilization options [35]

17
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Figure 10. Carbon capture options in SMR process [25]

Among all methods, absorption by MEA and MDEA are the industry standard and most
commonly used in CO2 removal [26], [35], [38]. MEA reacts rapidly with CO2 because of its high
alkalinity [38]. However, after the MEA has deactivated, regenerating it requires a large amount of energy
further increasing the cost of CCS technology [35], [38].

The capture rate of CO2 is in the range of 56% to 90% [26]. The captured CO2 is then compressed
and transported by pipelines or ships and stored in geological sites such as depleted oil and gas fields,
coal beds or stored in the ocean in saline aquifers at about or greater than 800 – 1000 m deep [39], [40].
At those depths, the CO2 is in a supercritical state and has a very high liquid-like density of around 500 –
800 kg/m3 [39]. This higher density would provide a higher packing efficiency for storing the CO2 with
annual leakage rates of about 0.00001 – 1% in geological sites and no subsurface leakage in
Saskatchewan, Canada’s enhanced oil recovery (EOR) and CCS site [37], [41].

Although, CCS has its positive effects in limiting the greenhouse gas output from SMR it does
have its downsides. With CCS technology added on to SMR, the total hydrogen plant and operating cost
will increase by 18% - 79% and 18% - 33% respectively [26]. The techno-economic evaluation by the
International Energy Agency (IEA) has a base cost to produce hydrogen at around USD $1.33/kg* from
SMR and an increase of about USD $0.24 – $0.60 if CCS is added [26]. There is some pushback in the
feasibility of blue hydrogen technology in lowering total GHG emissions [42]. But even without those
critiques, it is no secret that adding CCS technology would add to the total cost to produce hydrogen with
SMR and would not economically incentivize investors.

* - The costs were calculated using the values in the techno-economic assessment converted to USD
(using conversion rates on 09-29-2022) and a density of 0.08375kg/m3.

18
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

2.3.4 Electrolysis
Electrolysis is the endothermic electrification of water to dissociate H2O into its singular
components, hydrogen (H2) and oxygen (O2) as shown in equation (14). This process requires no carbon
containing feedstock and therefore carbon-free hydrogen. If the electricity used in the electrolysis process
originates from renewable sources, then this process of hydrogen production becomes self-sufficient
without reliance on GHG emitting sources. Water electrolysis from renewable energy is termed as
“green” hydrogen. This is the ideal form of hydrogen production since it releases little to no GHG, except
in initial capital costs to build a water electrolysis plant and necessary infrastructure to transport water and
electricity feedstock. The energy required to break down water at standard temperature (25°C) and
pressure (1 atm) is the enthalpy of formation for water of about 285.83 kJ/mol [12]. This is much higher
than the energy required for the SMR process of 253.03 kJ/mol in equation (9).

1
𝐻2 𝑂(𝑙) ↔ 𝑂2 + 𝐻2 , ∆𝐻° = 285.83 𝑘𝐽/𝑚𝑜𝑙 (14)
2

The first instance of dissociation of water through electrification is believed to date back to 1789
by van Troostwijk and Deinman [43]. The water electrolysis cell consists of two electrodes powered by
direct current and an ion-conducting electrolyte [44]. The induced electric field from the current causes
one electrode to possess a positive charge (anode) and the other to have a negative charge (cathode). The
direct current will produce ions at the surface of the electrodes where each electrode is termed based on
the type of ions attracted to it, positively charged ions (cations) attracted to the cathode and negatively
charged ions (anions) attracted to the anode [44]–[46]. The cations are attracted to the cathode because of
its negative charge from the electric field and respectively, the anions are attracted to the anode because
of its positive charge. Figure 11 shows a summary of this in addition to where the hydrogen and oxygen is
produced [47].

19
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

O2 Water H2

Anode Electrolyte Cathode

Current + – Electrons
Anions (–)
Cations (+)

Figure 11. Principles of an electrolyzer [47]

Equation (14) is a global equation that is broken down into half-cell redox reactions shown in
equation (15) and (16) or equation (17) and (18). The pH of the water electrolyzer plays a large role in
determining the makeup of the half-cell oxidation and reduction reactions due to the different ions it
produces. The reduction half-reaction takes places at the cathode where it possesses excess electrons from
the circuit thus attracting the cations. The oxidation half-reaction occurs at the anode where the circuit
causes electrons to leave the anode giving it a positive charge and attract the anions [46], [47].

In acidic aqueous solutions, the half-cell reactions of water electrolysis are [44]:

1
Anode: 𝐻2 𝑂 → 2 𝑂2 + 2𝐻 + + 2𝑒 − (15)

Cathode: 2𝐻 + + 2𝑒 − → 𝐻2 (16)

Equation (15) shows that in the anode, water will oxidize into oxygen, hydrogen ions and
electrons. Equation (16) shows that the hydrogen ions will reduce into hydrogen at the cathode [44], [46],
[47].

In alkaline aqueous solutions, the half-cell reactions of water electrolysis are [44]:

20
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

1
Anode: 2𝑂𝐻 − (𝑎𝑞) → 2 𝑂2 (𝑔) + 𝐻2 𝑂(𝑙) + 2𝑒 − (17)

Cathode: 2𝐻2 𝑂(𝑙) + 2𝑒 − → 𝐻2 (𝑔) + 2𝑂𝐻 − (𝑎𝑞) (18)

Equation (18) shows the water being reduced with the addition of electrons into hydrogen and
hydroxyl ions at the cathode. The hydroxyl ions at the anode will oxidize and form oxygen in equation
(17) [44], [46], [47]. As a result of these reactions, hydrogen will form at the cathode and oxygen will
form at the anode.

There are a few electrolysis technologies available mainly differing in the type of electrolyzer
being implemented. The most common water electrolysis technologies are alkaline water electrolysis
(AWE), proton exchange membrane electrolysis (PEM) and solid oxide electrolysis (SOEC) [47], [48].
Each technology has its host of advantages and disadvantages with respect to hydrogen production and
are further discussed in Appendix A.

2.3.5 Methane Pyrolysis


Methane pyrolysis (methane decomposition or methane cracking) is the decomposition of the
methane molecule (CH4) into its basic components: hydrogen (H2) and solid carbon (C) as shown in
equation (19). This is an endothermic reaction where the energy required to breakdown methane is its
respective enthalpy of formation of 74.87 kJ/mol [12]. Discussions about methane pyrolysis usually come
with comparisons against its dominant counterpart for hydrogen production, SMR. In methane pyrolysis,
equation (19), one mole of methane produces two moles of hydrogen. When compared to SMR, equation
(9), one mole of methane releases four moles of hydrogen. In methane pyrolysis, two times as much
methane is required to produce the same amount of hydrogen in SMR. When looking at Table 3, the
energy required to produce one kilogram of hydrogen from methane decomposition (18.72 MJ/kg) is
lower than that of SMR (31.63 MJ/kg). The energy requirement for SMR includes the energy required to
heat liquid water into water vapor. It should be noted that the values in Table 3 are the minimum
theoretical values required for the reaction to occur. Actual values will be higher when considering
efficiencies of the overall system and its life cycle.

𝐶𝐻4 → 𝐶(𝑠) + 2𝐻2 (𝑔), ∆𝐻° = 74.87 𝑘𝐽/𝑚𝑜𝑙 (19)

21
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Table 3. Energy required to produce hydrogen in SMR, electrolysis and methane pyrolysis

Hydrogen Production Energy per mole of H2 (kJ/mol) Energy per kilogram of H2


Type (MJ/kg)

Steam Methane 63.258 31.63


Reforming

Electrolysis 285.83 142.92

Methane Pyrolysis 37.435 18.72

Achieving moderate rates of non-catalytic methane decomposition don’t occur until high
temperatures of at least 1000°C due to high activation energy requirements. The activation energy
required for non-catalytic methane pyrolysis varies in the literature between 356 kJ/mol – 452 kJ/mol
[49].

On top of the hydrogen, methane pyrolysis also produces carbon as a by-product. By weight,
there is three times more carbon produced than hydrogen. Depending on the form of the resulting carbon,
if a suitable application can be found for it, it can become an additional source of income. At a worse case
scenario, if the carbon has no use case it can be sequestered underground. Storing solid carbon
underground is far easier to do compared to producing blue hydrogen and attempting to sequester gaseous
CO2. This method of hydrogen production has been termed as “turquoise” hydrogen.

In reality, similar to SMR, natural gas would be used as the methane source. If inexpensive
natural gas can be provided and the energy used to crack methane is renewable, methane pyrolysis can
produce hydrogen with little to no CO2 emissions with energy economics comparable to SMR and
potentially marketable carbon by-product (if not, it can be sequestered). These positive attributes make
methane pyrolysis quite attractive for large-scale and economic hydrogen production while
simultaneously serving as a bridge to the hydrogen economy [7]. However, this technology possesses a
low technology readiness level (TRL) since it is mainly in the R&D phase.

Current research involves discovering the correct reaction mechanism for carbon formation along
with the most optimal reactor type and catalyst [50]. Multiple reactor types have been proposed and
studied such as plasma, a packed bed, fluidized bed, monolithic, liquid bubble column or a moving bed
reactor [25]. Figure 12 shows the different reactor designs with its respective heating source and operating
temperature range [51]. It should be reiterated that the proposed design of study in this thesis, fluidized
bed reactor in a microwave oven, is not listed in the figure indicating the lack of research in this design.

22
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Figure 12. Reactor types with heat source and operating temperature range [51]

Since different designs require different operating temperatures, especially in the presence of a
catalyst, this raises a question on the type of carbon produced and its resulting morphology. Keipi has
produced an excellent image (see Figure 13) summarizing the literature that reported the different types of
carbon produced in its respective reactor design [51]. Different metals in liquid metal reactors produce
different types of carbon such as graphite-like carbon, carbon nanotubes, carbon nanofibers. Filamentous
carbon products appear to be more common among metal-based catalysts where carbon catalysts or non-
catalytic methane decomposition tends to output graphite-like carbon or carbon black [51]. Reactors in
very high temperature regimes of at least 1300°C, such as plasma reactors, also tend to produce carbon
black [52]. The main problem in most reactors is clogging of the reactor over time due to carbon
deposition. Methods to circumvent this issue need to be implemented for continuous hydrogen
production.

23
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Figure 13. Types of carbon produced expressed as a function of types of catalyst and temperature [51]

Methane pyrolysis can be divided into three different categories: thermal decomposition, plasma
decomposition and catalytic decomposition [53]. Each of these respective forms of methane
decomposition can be further divided into its respective sub-categories as outlined in Figure 14 [53].

24
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Reactor walls

Thermal Liquid metal


Decomposition

Carbon
granules

Plasma Plasma torch


Decomposition

Metal catalyst
& plasma torch

Catalytic Metal catalyst


Decomposition

Carbon catalyst

Figure 14. Categories of methane pyrolysis [53]

Thermal decomposition is the breaking down of methane in temperature regimes more than
1000°C where the heat source is from the reactor walls, hot liquid metal or carbon granules [5]. Plasma
decomposition involves exciting and ionizing the gas stream with localized temperatures of 2000°C using
a plasma torch [53]. This provides more efficient heat transfer that accelerate the reaction [54]. Catalytic
decomposition involves methane cracking at more moderate temperatures but involves a metallic or
carbonaceous catalyst to accelerate the reaction. These sub-categories of methane pyrolysis will be further
explored in the following sections below.

2.3.5.1 Thermal Decomposition


As previously mentioned, thermal decomposition is the breakdown of methane at elevated
temperatures without the presence of a catalyst to accelerate the reaction. As the temperature increases,
the methane to hydrogen conversion rate increases as well [55], [56]. The reactor designs that have been
studied in this category are direct heating by a furnace [56], liquid metal baths [57]–[64], solar radiation
[65], carbon granules in a packed bed reactor (PBR) [66], [67] or fluidized bed reactor (FBR) [68], [69].

25
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

In a liquid metal bath, there is a column of liquid metal of temperatures between 700°C – 1500°C
[57], [60]. The composition of the liquid metal bath depends on the design but typically consists of alloys
of the following metals, but not limited to, tin (Sn), lead (Pb), copper (Cu), nickel (Ni), bismuth (Bi) or
tellurium (Te) [57], [60], [61], [64], [70]. The methane will flow from the bottom through the liquid metal
solution where the reaction in equation (19) will take place. The hydrogen will bubble out and the carbon
will float to the top (see Figure 15) [61], [64]. Some of the metals evaporate under high temperatures and
can condense into the pores of the carbon. The carbon is therefore contaminated with the metal in the bath
ranging from about 0.7 wt.% tin contamination in carbon for a pure molten tin bath [58] compared to
about 83 wt.% metal contamination in carbon for a Ni-Bi bath [62].

Figure 15. Schematic of methane pyrolysis in a liquid metal bath [64]

In the case of a reactor design containing carbon granules, the carbon granules are heated up,
typically by wall heating or in a furnace to suitable temperatures where methane will pass through it
where pyrolysis can take place (see Figure 16). The carbon granules can be a packed bed, but this presents
a slight problem. After the pyrolysis reaction in equation (19), the carbon from the methane will deposit
on the carbon granules, increasing its size to an extent where it will cause blockage in the flow area [68].

Muradov evaluated various reactor types such as a packed bed reactor, fluidized bed reactor, free-
volume reactor, fluid wall reactor and a tubular reactor and concluded that a fluidized bed is more suitable
to work around the blockage issue [66]. The idea is that in a fluidized bed, the particles move around and
allow more space for particle growth. This prevents blockage over longer periods thus making it more
suitable for operations requiring large-scale hydrogen production [66], [71]. Under fluidization, the
carbon particle bed will behave as a liquid and assist in the carbon particle-to-gas heat transfer rate and
provide a more uniform temperature distribution. The challenge associated with a fluidized bed reactor is
that the particles will still grow which may affect the fluidization and cause blockage. A procedure to
remove the larger particles will need to be implemented in order for the process to be optimally operated.

26
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Figure 16. Schematic of methane pyrolysis in a fluidized bed reactor composed of carbon granules [66]

The methane to hydrogen conversion rate depends on the reaction conditions such as temperature,
residence time and the presence of a catalyst [72]. The non-catalytic methane to hydrogen conversion rate
in a furnace is around 74.6% at 1350°C to about 99.7% at 1700°C [56].

Some metals and carbon granule-based reactors can behave as catalysts and assist the pyrolysis
reaction to occur at lower temperatures [73]. For the reaction to happen due to thermal effects rather than
a catalytic effect, the catalyst itself needs to be deactivated. More information on catalytic decomposition
and catalyst deactivation can be found in section 2.3.5.3 Catalytic Decomposition.

2.3.5.2 Plasma Decomposition


Plasma decomposition is the introduction of a gas stream into an electrically generated plasma at
temperatures in the range of 1200°C – 2200°C to produce a continuous stream of hydrogen [53], [74],
[75]. Plasma process’s main goal is to economically produce carbon black with hydrogen as a secondary
product without any CO2 emissions [52], [74], [76]–[78]. Monolith is a company that uses a DC plasma
torch to make hydrogen and carbon black [78], [79]. The advantages of the plasma process allow the near
100% conversion of methane to hydrogen and carbon black [76]. In addition, it has the capabilities of
being able to also produce other high quality carbon products such as carbon nanotubes and graphene
nanosheets [77] where its properties and quality can be modified depending on the temperature of the
system [75].

A typical plasma-based pyrolysis reactor is shown in Figure 17 [76]. Graphite electrodes at the
top of the reactor receive power from a power supply that will heat up the reaction chamber where

27
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

hydrocarbons are injected. The pyrolysis reaction occurs, and the exhaust is sent to a filter where the
carbon black is separated from the gases.

Figure 17. Schematic of plasma pyrolysis reactor [76]

By-products of plasma methane pyrolysis include ethane, ethylene, acetylene and


benzene [51], [52]. In a computational model, temperatures higher than 2200°C results in an increase of
ethylene and acetylene due to a gradual decomposition of hydrogen molecules into hydrogen atoms [80].
The main disadvantage is that the plasma process requires large amounts of energy to achieve the
temperatures in the plasma state necessary for complete methane decomposition.

2.3.5.3 Catalytic Decomposition


Since methane pyrolysis involves temperatures of at least 1000°C to achieve moderate hydrogen
output, catalysts are utilized to help facilitate comparable results but at a lower temperature. Metal and
carbon are the most extensively studied and common catalysts used. Photochemical catalysts have also
been studied but this suffers from low methane conversion and poor hydrogen selectivity [81]. The
following sub-sections provide more details on metal and carbon catalysts.

28
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

2.3.5.3.1 Metal Catalysts

Metal catalysts are typically used as a liquid in molten metal reactors of temperatures around 600
– 900°C [82]. Some metals that have been used in molten metal pyrolysis are, but not limited to, copper
(Cu) [64], [83], tin (Sn) [57], [59], [63], [83], platinum (Pt) [84], lead (Pb) [59] or iron (Fe) [85]. Nickel-
based catalysts are the most commonly used and successful in methane conversion with high H2 yield at
moderate temperatures due to their high catalytic activity [86]. Upham [61] has mentioned that the
problem with active metals (e.g. Ni, Pt, or Pd) that are good for catalytic methane decomposition is that
they require a high melting point temperature (~1400°C for Ni) that is much higher than the temperature
needed for methane pyrolysis. Low melting temperature metals, such as Sn, Pb, In or Ga, don’t have
strong catalytic effects. To produce active catalysts that melt in the range of 600°C – 900°C, alloys of
these active metals are dissolved in the low melting point solvents (Sn, Pb, Bi, In and/or Ga) [61]. Some
alloys that have been made based on this is a nickel-bismuth catalyst that achieved 95% methane
conversion at 1065°C [61] or a copper-bismuth catalyst [64].

The main advantage of the catalytic molten metal bath is the efficient heat transfer from the hot
metal to the methane stream and the ease of carbon collection as it floats to the top [61], [64], [87].
However, the carbon is contaminated with metal with about 83 wt.% metal contamination of carbon in a
Ni-Bi reactor before any purification [62]. This issue limits the application of where the carbon can be
used if such contaminants remain in the system.

Another issue with catalytic methane pyrolysis is the deactivation of the catalyst due to carbon
deposition (coking) on active sites [5], [61], [88], [89]. Catalyst deactivation leads to a decrease in the
methane to hydrogen conversion rates at the same temperature [59], [84]. The process will be required to
be shut down to change the catalyst, or another process could be implemented to regenerate the catalyst.
The most common methods for catalyst regeneration are to burn off the carbon covering metal by air or
steam regeneration. There are some tests done showing that steam regeneration is capable of completely
regenerating a metal catalyst without significant loss of catalytic activity [90], [91]. However, catalyst
regeneration defeats the purpose of low-CO2 hydrogen production as the regeneration process will
produce carbon emissions and provide further complexities and cost to the process.

2.3.5.3.2 Carbon Catalysts

Carbon catalysts have been studied in packed bed or fluidized bed reactors instead of metal
catalysts to get around the issue of metal contamination in carbon and catalyst deactivation in metal
catalysts. Some types of carbon catalysts that have been studied are carbon black, activated carbon,

29
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

graphite, carbon fiber, glassy carbon, synthetic diamond powder, acetylene black, carbon nanotubes, soot
(fullerenes) [92], [93]. Advantages carbon catalysts offer over metal catalysts are, but not limited to [5],
[87], [94]–[96]:

• Carbon catalysts are resistant to high temperatures


• Resistant to sulfur impurities
• Low cost
• Widely available
• No contamination of catalyst with carbon product
• No requirement for carbon separation from catalyst unlike metal catalysts
• Produced carbon may have autocatalytic effects
• No need for catalyst regeneration

The caveat to these advantages is that the required operating temperature for carbon-based
catalysts are higher than metal-based catalysts because of a higher activation energy requirement in the
carbon case [97].

It is believed that catalytic activity of carbons is determined by their origin, structure and surface
area [92], [93]. Muradov has experimented with a variety of carbon materials with different crystalline
structures and surface morphology to observe their catalytic effect. He has found that efficient methane
pyrolysis can be achieved on high surface area amorphous carbons such as activated carbon and carbon
black [87], [92], [93]. Amorphous carbons tend to have disrupted arrays of carbon bonds that form free
valences, discontinuities, edges, defects and other abnormalities that have been described as high-energy
sites (HES) [93]. The more HES there are, the higher the initial activity of the carbon catalyst. Activated
carbon and carbon black tend to have a very high amount of surface area and therefore a large amount of
HES. It is believed that the mechanism behind methane pyrolysis over carbon catalysts is initiated with
the adsorption of methane molecule followed by multiple chronological dissociation reactions to satisfy
their valency requirements and stabilize energetically [93]. Therefore, the order of activity in carbon
catalysts is determined to be amorphous carbon (disordered) > turbostratic (less ordered) > graphite
(highly ordered) [93].

The activation energy of activated carbon and carbon black is 160 – 201 kJ/mol and 205 – 236
kJ/mol respectively [98]. Generally, activated carbon catalysts are initially more active compared to
carbon black catalysts (assuming similar surface area) but the catalytic activity in activated carbon
catalysts diminish at a faster rate than carbon black catalysts [93], [95], [98].

30
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Similar to metal catalysts, carbon catalysts are prone to deactivation as well due to carbon from
methane decomposition blocking its pores and lowering its surface area [68], [87], [92], [93], [98].
Although, the catalytic activity drops from prolonged carbon deposition, the methane conversion
stabilizes possibly due to the product carbon possessing some catalytic effect of its own (autocatalysis)
[69], [99]. The carbon catalyst could be regenerated through steam or air treatment with success in
completely restoring its original catalytic activity [98]. Although an autocatalytic effect from the
produced carbon would be ideal, catalyst regeneration could be employed in continuous hydrogen and
carbon production in a fluidized bed reactor. However, this would produce CO2 emissions.

2.4 Microwave Radiation


The problem with conventional heating methods (such as wall heating) is that they are inefficient.
In non-plasma methane pyrolysis, conventional heating provides a temperature gradient occurring radially
from the walls of a reactor to the center which limits methane conversion rates [100]. Uniform
temperatures may be too slow to achieve in applications that require larger diameter reactors. Large
applications also tend to demand minimized costs and maximized heat transfer and efficiencies. One
method of providing thermal energy that is more efficient than conventional methods are microwave
heating which have seen some success in methane pyrolysis [100].

Microwaves are a form of radiation on the electromagnetic spectrum (see Figure 18) that move at
the speed of light [101]. The wavelengths of microwaves range from 1 m to 1 mm corresponding to a
frequency range of 300 MHz to 300 GHz. Rather than conventional heat transfer, microwave heating
involves energy conversion from electromagnetic energy to thermal energy. Unlike conventional heat
transfer where it’s driven by thermal gradients, microwave heating is applied directly to the medium
through molecular interaction with the electromagnetic field resulting in volumetric heating [9]. This
method of delivering contactless energy is more efficient than conventional heat generation methods [8]
since it can generate heat where it is required [9]. Microwave heating offers several other advantages such
as [9]:

• Volumetric heating
• Selective material heating
• Rapid heating
• Non-contact heating
• Quick turning ON and OFF
• Ability to treat waste in-situ
• Portability of equipment and processes

31
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

These advantages have allowed the exploration of applications of microwave heating in many
fields such as food processing, medical waste treatment, contaminated soil remediation, wastewater
cleanup, activated carbon regeneration, pollution control, pyrolysis [9].

Figure 18. Electromagnetic spectrum [101]

Microwave radiation interacts with materials in three different ways: as a microwave-transparent


material (quartz), conductors that reflect microwaves (metals) or microwave absorbers (water, carbon).

The transparency of a material in a microwave field is dependant on its relative permittivity. A


low relative permittivity means that it will not absorb any microwaves. Quartz has a very low relative
permittivity which is why it’s transparent to microwave energy and commonly used in microwave
systems.

In microwave heating (also known as dielectric heating), water heats up due to interactions with
the electric field. The O-H bonds in water molecules are polar covalent bonds that oscillate rapidly under
an electric field. These oscillations cause the molecules to heat up rapidly thus heating up the entire
volume of the medium. The variable associated with dielectric heating is the dielectric loss factor. The

32
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

dielectric loss factor is the ability of a material to absorb energy when an electromagnetic field is shone at
it. The higher the dielectric loss factor, the more energy the medium will absorb.

Carbon containing materials are well known to be good microwave absorbers [102]. Specifically,
activated carbon is also a suitable microwave receptor as it’s able to reach temperatures over 900°C in
over a minute when irradiated with 1000 W [100], [103]. However, carbon isn’t a polar molecule so it
can’t be heated up by the electric field. The carbon is an electrically conductive material that is heated by
induction heating from the magnetic field in microwave energy. This is caused by the magnetic field
inducing eddy currents within the structure of the carbon. The induced eddy currents will heat up the
carbon by joule heating (also known as ohmic heating or resistance heating).

The penetration depth of microwaves is determined by the frequency of the microwave and the
dielectric loss and magnetic loss factors of the microwave medium [9]. The penetration depth of a
microwave decreases with an increase in frequency [9]. Typically, the frequency of a microwave is set so
the only factor that influences the microwave penetration depth is the dielectric loss factor associated with
the material. Generally, the higher the dielectric loss factor, the lower the penetration depth (and vice
versa) [9]. To maximize microwave absorption, the material should be of a size that is comparable to the
penetration depth.

A magnetron is used to generate microwave radiation when a large potential difference is applied
to the anode and cathode inside the magnetron. This potential difference ejects electrons which are then
guided by magnetic fields causing the electrons to travel in a spiral direction (also known as an electron
beam). As electrons travel through the resonant cavity, oscillations are built in the electron cloud where
its frequency is dependent on the cavity size. Since cavity size is generally fixed, the frequency emitted in
magnetron tubes are also fixed to one specific value. The output power of the magnetron can be
controlled by adjusting the current or magnetic field strength [104].

2.5 Fluidization
Fluidization consists of an operation where granular particles transition from a solid behaving
state to a fluid like state when a fluid is passed upwards through the particles. In a stationary bed of
particles, if a fluid is slowly passed upward through a bed of particles, the fluid passes around the spaces
in between the particles without disturbing the particles’ current position and orientation. If the flow rate
were to increase, the particles start to move apart due to a produced drag force in the direction of the flow.
This is known as an expanded bed. If the flow velocity is increased even further, the particles reach a state
where the flow suspends the particles upward due to an increased drag force where it will balance out

33
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

with the gravitational force of the particles. This is termed as a fluidized bed. Fluidization begins at a
velocity where the gravitational force on the carbon particles is equal or less than the drag on the particles
causing the bed to expand and the void fraction to increase. There are many fluidization regimes which
are dependant on the flow velocity and can be seen in Figure 19 below [10].

Figure 19. Fluidization regimes [10]

Fluidized bed reactors offer many advantages that a fixed or packed bed reactor doesn’t, such as,
but not limited to [10], [105]:

• Excellent heat and mass transfer rates between gas and particles when compared to other modes
of contact heat transfer
• Uniform temperature distribution due to particle and gas mixing
• Accommodating to particle growth
• Suitable in large scale operations

34
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

Some of its disadvantages are:

• Non-uniform gas distribution due to bubbling or turbulent flows


• Intensive back mixing of solid particles that reduced performance for catalysts attempting to
capture and release reactant gas molecules
• Erosion of reactor due to particle abrasion
• Attrition of particles that may affect fluidization
• Particle growth that may affect fluidization

Some applications of gas-solid fluidized bed reactors are in heat exchange, particle growth, coal
gasification or even pyrolysis of biomass [105].

The minimum gas velocity required for fluidization is termed as the minimum fluidization
velocity, umf. The minimum gas velocity required to observe the formation of bubbles is termed as the
minimum bubbling gas velocity, umb. The fluidization of gas-solid systems is strongly dependent on the
types of particles which are classified based on their fluidization behavior and size as seen in Figure 20
[106].

Figure 20. Geldart's classification of fluidized particles [106]

Group A particles are of size 30 – 100 micrometers and are easy to fluidize. The minimum
fluidization velocity (umf) is less than the minimum bubbling velocity (umb) for particles in group A.
Group B particles are greater than 100 micrometers and are also easy to fluidize and form bubbles. The
minimum fluidization velocity is similar to that of the minimum bubbling velocity. Group C is too small

35
Chapter 2: Hydrogen in the Energy Landscape and the Influencing Factors of Methane Pyrolysis

and difficult to fluidize due to interparticle forces effecting fluidization such as van der Walls and
electrostatic forces. Group D is too big to fluidize as it requires a higher fluidization gas velocity to
achieve a fluidization state. Particle sizes of 100 – 500 micrometers seem to be optimal for fluidized
methane cracking [66].

The minimum fluidization velocity can be determined by first calculating the voidance fraction,
εmf, between the particles at minimum fluidization as per equation (20) below [105].

0.029
𝜇2 𝜌𝑔 0.021
𝜀𝑚𝑓 = 0.586𝜙𝑠−0.72 ( 3
) ( ) (20)
𝜌𝑔 (𝜌𝑠 − 𝜌𝑔 )𝑔𝑑𝑝 𝜌𝑠

where 𝜀𝑚𝑓 is the voidance fraction, 𝜙𝑠 is the particle sphericity (from 0 to 1), 𝜇 is the dynamic viscosity,
𝜌𝑔 is the gas density, 𝜌𝑠 is the particle density, g is the gravitational constant and dp is the particle
diameter.

The minimum fluidization velocity equation is dependent on the Reynolds number of the particles which
is calculated by equation (21).

𝜌𝑔 𝑑𝑝 𝑢
𝑅𝑒𝑝 = (21)
𝜇

If Rep < 10, minimum fluidization velocity is calculated by equation (22).

2 3
(𝜙𝑠 𝑑𝑝 ) (𝜌𝑠 − 𝜌𝑔 )𝑔 𝜀𝑚𝑓
𝑢𝑚𝑓 = ( ) (22)
150 𝜇 1 − 𝜀𝑚𝑓

If Rep > 1000, the minimum fluidization velocity is expressed as per equation (23).

3 1/2
𝜙𝑠 𝑑𝑝 (𝜌𝑠 − 𝜌𝑔 )𝑔𝜀𝑚𝑓
𝑢𝑚𝑓 =[ ] (23)
1.75 𝜌𝑔

36
Chapter 3: Experimental Apparatus

Chapter 3: Experimental Apparatus


This chapter consists of the description of the experimental apparatus involving the microwave
unit apparatus, fluidization, carbon particles and exhaust analysis. Each of these described sections will
discuss how every design aspect of microwave methane pyrolysis is integrated and its functions. The
appropriate operational instructions and calibration steps can be found in the appendix.

3.1 Microwave Unit Preparation


This section involves the discussion of every part and instrument necessary for the successful
operation of microwave methane pyrolysis. Figure 21 shows the skeletal outline of the experimental
process (from A and view in a clockwise direction). Each of the following listed instruments will be
discussed in more detail in their own respective sub-section.
A – Methane and nitrogen cylinders. The methane is diluted with nitrogen for safety at around 50%
methane to 50% nitrogen ratio.

B – Microwave Chokes. They prevent microwaves from escaping the microwave cavity while also
allowing the attachment of the quartz reactor into microwave.

C – Fritted discs. These porous discs hold the carbon particles in place and allow gas to pass through.

D – Microwave Unit. A 1 kW, variable power level at 2.450 GHz.

E – Thermocouple. Temperature is measured from the top where LabVIEW reads it.

F – Muffle. The muffle helps insulate the heat in the reactor vessel.

G – Quartz reactor filled with carbon particles. A 28-inch-long quartz tube with 25mm OD and 22mm ID
is used as the reactor vessel that contains the carbon particles.

H – Gas chromatography – thermal conductivity detector (GC-TCD). The GC unit analyzes the exhaust
gas for its composition.

37
Chapter 3: Experimental Apparatus

Figure 21. Schematic of experimental setup for microwave-driven methane pyrolysis: (A) Methane and nitrogen cylinders, (B)
microwave chokes, (C) fritted discs, (D) 1000 W microwave, (E) thermocouple, (F) muffle, (G) reactor filled with carbon
particles, and (H) GC – TCD

With respect to Figure 21, the basic operation of this process is that a mixture of 50% methane
and 50% nitrogen (A) will flow from the bottom through the chokes (B) and fritted discs (C). The
microwave (D) will turn on and heat up the particles (212 – 425 micrometers sized particles) in the quartz
reactor (G) which is insulated by a muffle (F). The incoming flow will fluidize the carbon granules and
assist in heat transfer between the high temperature carbon particles and the gas itself. The temperature is
measured by a k-type thermocouple (E) in a singular location in the middle of the reactor. The exhaust
gases flow towards the GC-TCD (H) where its output composition will be analyzed. Figure 22 below
shows the actual setup.

38
Chapter 3: Experimental Apparatus

Figure 22. Experimental setup for microwave-driven methane pyrolysis: (A) Vent, (B) H2 sensor, (C) thermocouple, (D) exhaust
line, (E) microwave, (F) inlet line, (G) pressure gauge, (H) pressure relief valve

The setup and operation instructions of the experimental process can be found in Appendix B and
C respectively.

39
Chapter 3: Experimental Apparatus

3.1.1 Microwave Unit


The microwave used in this experiment is a 1 kW with variable power level at 2.450 GHz unit
that requires 210 V, 15 A and 60 Hertz. It is a unit with a cavity size of 13” (330mm) x 13” (330mm) x 8
1/16” (204.7mm) tall. Is has been modified with two holes on the side to allow metal chokes to fit through
and accommodate the reactor. It has cooling fans mounted in certain areas of the microwave to cool the
electronics. It also has a variable power level knob (from 0 to 100) to control the power. The numbers on
the knob do not correlate to the percentage of maximum power the microwave outputs. The numbers are
valueless, and it is not known how much microwaves are emitted at a given power level setting. The
model is a BP-095 purchased from Microwave Research & Laboratory Inc. and shown in Figure 23.

Figure 23. 1 kW, 2.450 GHz microwave unit

40
Chapter 3: Experimental Apparatus

3.1.2 Microwave Chokes


The microwave chokes are two stainless steel cylinders threaded at one end so that they may be
attached to each end of the microwave see Figure 24. The image on the left shows the choke attached and
the image on the right is how it looks outside of the microwave. When attached, the chokes stick out on
the sides of the microwave in such a way where a 1-inch reactor can slide inside the chokes all the way
through from end-to-end. They have been designed in such a way such that when the microwave is turned
on, no microwaves leak from the cavity.

Figure 24. Microwave chokes attached to microwave (left) and detached (right)

41
Chapter 3: Experimental Apparatus

3.1.3 Reactor Vessel


The reactor vessel is a 28-inch length fused quartz tube of 25 mm outer diameter (OD) and 22
mm inner diameter (ID) purchased from Technical Glass Products, Inc. 11 inches from one end, the
quartz has been choked to appear more as an hourglass as shown in Figure 25. Note that it is not a dimple
but a continuous indent at the 11-inch mark. The continuous indent is small enough so that a 20 mm
fritted disc can sit on top of that area allowing for the carbon sample to sit on the discs. The quartz is
transparent to microwaves which is necessary to allow the microwaves to enter the reactor and penetrate
the carbon particles inside for it to be heated. When considering different reactor materials, transparency
to microwaves must be taken into consideration. The strain, annealing and softening temperature of the
quartz are 1120°C, 1215°C and 1683°C respectively. Other properties of the quartz tube can be found on
the manufacturer’s website at the time of writing of this thesis or in Appendix E. Note that quartz and
reaction vessel are interchangeably used throughout this thesis as both terms refer to the glass tubing
shown in Figure 25.

A 25.4 mm stainless steel disc with a 40-micron pore size was placed near the outlet to prevent
the particles from flying out of the reactor.

Figure 25. Quartz reaction vessel

42
Chapter 3: Experimental Apparatus

3.1.4 Fritted Disc


The fritted quartz discs are white 20mm diameter porous discs with a 3 – 5mm thickness as
shown in Figure 26. The porosity of these discs is around 40 – 100 micrometers which is smaller than the
size of the carbon particles (212 – 425 microns). This allows the carbon particles to sit on the discs
without clogging its pore while simultaneously allowing gas to pass through. Three discs are typically
placed in the reactor for stability. Any less would cause the discs to tilt over thus disabling the ability for
the carbon to sit on top of the discs.

Figure 26. Fritted discs 20 mm in diameter

43
Chapter 3: Experimental Apparatus

3.1.5 Flow Meters


Two different flow controllers are used for this apparatus. One flow controller directs the nitrogen
(Gr4.8 300SZ) and the other controls the flow rate of methane (Gr2.0 200SZ). The methane flow is
controlled by a Brooks flow controller capable of controlling flow rates of up to 0.3 L/min at standard
conditions. The Bronkhorst controller capable of controlling flows up to 4 L/min controls the nitrogen
flow. The flow rates are set to 0.1 L/min for each controller, nitrogen and methane. The flow lines are
combined into one flow line (union tee) so that the total flow rate would be the summation of the flow
rate in each line. Since each flow line is ran at 0.1 L/min, this would give a total flow rate of 0.2 L/min
and consequently 50% methane and 50% nitrogen. The total flow rate for most experiments is maintained
at 0.2 L/min for fluidization purposes and the concentrations of methane and nitrogen are adjusted by
changing the flow rates in their respective line (see section 3.2 Fluidization for more information on
fluidization).

The flow controllers are calibrated for the respective gas being used. The most recent calibration
curves for each gas and are shown below in equations (24) and (25). The operational instructions for flow
meters and the calibration procedure can be found in Appendix C and F, respectively.

Nitrogen:

𝑦 = 2.18𝑥 + 0.009 (24)

where x is the value input into the controller and y is the real flow rate in L/min.

Methane:

𝑦 = 0.000827𝑥 − 0.00625 (25)

where x is the value input into the controller and y is the real flow rate in L/min.

The outlet of the union tee is attached to a 3-meter-long looped nylon hose connected to a
pressure relief valve and pressure gauge before the reactor inlet.

44
Chapter 3: Experimental Apparatus

3.1.6 Insulation
When the reactor vessel is heated up from the microwaves, the heat loss to the surroundings can
be very large and be a hindrance when attempting to obtain temperatures of around 1000°C. To mitigate
heat loss, microwave transparent insulation is placed inside the microwave. The insulation is a muffle
made of alumina (Al2O3) and quartz (SiO2) rated for about 1200°C continuous duty. The muffle is 10.5-
inch x 1.5-inch x 1.5-inch that can house a 1-inch reactor vessel inside the microwave cavity and has a
door on one end to allow for viewing if desired as shown in Figure 27. For maximum insulation, the door
is kept on most of the time to prevent heat loss and the space in between the reactor and the interior
muffle walls are stuffed with more insulation material for additional heat containment. Achieving high
temperatures has been quite a challenge in tests surrounding this experimental apparatus. As much
insulation as possible was placed inside the microwave cavity to contain the heat and minimize any losses
to the environment. This step has been extremely useful in obtaining temperatures reaching the 1300°C
regime.

Figure 27. Muffle insulation with no door (left) and door (right)

45
Chapter 3: Experimental Apparatus

3.1.7 Sealing
Some areas of the microwave experimental unit that need to be sealed are the inlet, outlet and its
corresponding flow lines. Sealing near the inlet and outlet needs to be proceeded with caution since the
reactor vessel is made of quartz and quite fragile. The inlet and outlet are sealed using O-rings that
squeeze onto the quartz vessel preventing any inlet or outlet gases from escaping. Since the inlet and
outlet plumbing do not get very hot, the O-rings do not melt and are able to withstand the modest amount
of heat escaping from the ends of the unit. The ¼ inch flow lines are connected by Swagelok connections
and ferrules where the threads have been wrapped in sealing tape to seal the flow lines and prevent gas
leaks. When the larger plumbing (1 inch) is connected to the ultra torr connection, the heat from the
friction can weld the threads together. To prevent this while also sealing it, anti-seize and sealing paste
instead of sealing tape are used. The anti-seize is used to prevent the 1-inch connections from galling,
seizing and corrosion, as well as lubricating to ease assembly. The sealing paste is used to seal the larger
connectors to each other rather than the tape.

Figure 28. Ultra torr sealing Swagelok connection

3.1.8 Temperature Sensor


A k-type Super OMEGACLAD® XL thermocouple (purchased from OMEGA) with 6.35 mm
diameter, rated for a temperature of up 1350°C, is used to measure the temperature of the system. It is
placed in the middle of the reactor and inserted from the top as shown in Figure 22. Due to the design of
this specific setup, only one thermocouple can be inserted in the reactor. The thermocouple is connected
to a computer where the local temperature is continuously displayed.

46
Chapter 3: Experimental Apparatus

3.1.9 Ventilation
The entire experimental apparatus unit is placed vertically on a stand where a large plexiglass box
encloses the setup. As can be seen in Figure 29, a square cut out is at the top where a fan is placed
sucking out the ambient air inside the box leading to a fume hood. This is so that if any leakage were to
occur, it would be safely extracted and dumped into the fume hood. An additional feature of the fan is that
it vacuums the immediate warmer air providing circulation and cooling the room itself.

Figure 29. Ventilation fan

3.1.10 Safety
Several safety layers are employed to reduce the chance of microwave leakage and any explosion.
A microwave leak detector was utilized to check the leakage around the microwave. Whenever the leak
exceeded 5 mW/cm2, the nut that tightens the choke was covered with multilayers of aluminum foil to
minimize microwave leakage. Another potential safety issue is combustion of methane and hydrogen. To
mitigate this, the methane is usually run at 50% concentration with the other half being nitrogen. This is
to ensure a low concentration of hydrogen in the output in case of a leak. If a hydrogen or methane leak
were to occur, the system is enclosed in a plexiglass chamber with a fan venting the immediate air into a
fume hood. In the event of a hydrogen leakage, the Honeywell BW Solo Single-Gas Detector hydrogen
sensor placed near the fan in the top would detect the hydrogen from 0 to 1000 ppm. If the hydrogen or
methane leak were to combust inside the microwave or the box, there are layers of protection. If a
methane/hydrogen mixture combusted inside the reactor, the muffle is there to contain it, followed by the
microwave cavity and the plexiglass. In the event of a fire, the methane line would be immediately shut
off with two fire extinguishers in the lab on standby.
47
Chapter 3: Experimental Apparatus

3.2 Fluidization
3.2.1 Particle Size
As section 2.5 Fluidization outlines, there are quite a few fluidization regimes that are dictated by
the flow velocity and the size of the particles in the reactor as well. Particle sizes of 100 – 500
micrometers seem to be optimal for fluidized methane cracking [66]. For the experiments, particles sizes
of 212 – 425 micrometers are used as the base particles for fluidization.

3.2.2 Fluidization Regimes


The main fluidization regimes of interest are minimum fluidization, bubbling fluidization,
slugging flow and turbulent fluidization. It was found experimentally that minimum or bubbling
fluidization is the best regime to be in for effective heat transfer and achievement of high temperature. If
the velocity of the flow was any higher, therefore higher fluidization regime, then the carbon particle
medium would be cooled due to convective heat transfer as the gas is coming in at room temperature. The
higher velocity also results in a smaller residence time for the reaction to take place.

It should be noted that as the gas temperature gets hot upon interaction with the hot carbon
particles, the density of the gas changes due to the heating effect. This will affect the flow velocity of the
incoming gas and affect fluidization. To combat this, the minimum fluidization velocity was
experimentally obtained by trial and error where the flow rate was adjusted at around 1200°C until the
fluidization regime visually changed from slugging to bubbling fluidization. The flow rate that was
determined from this and been used for all experiments is 0.2 L/min.

48
Chapter 3: Experimental Apparatus

3.3 Carbon Particles


The nature of the carbon particles is important as it should be small enough for fluidization and be
able to absorb microwaves as its main properties. This is further described in the following sections.

3.3.1 Carbon Particle Properties


It has been speculated that in methane pyrolysis involving carbonaceous particles, surface area is
a major factor in causing methane-to-hydrogen conversion [68]. Activated carbon (Matrix Carbon)
purchased from Seachem Laboratories, Inc. was used as the initial particles inside a fluidized bed reactor
to absorb microwave energy. Activated carbon is very porous possessing large amounts of surface area
and is historically used as a catalyst in catalytic methane pyrolysis [93]. Activated carbon is known to be
a good absorber of microwaves making it a very good substance in microwave methane pyrolysis in a
fluidized bed [107].

3.3.2 Grinding Particles


The carbon particles are initially received much larger than its intended size of 100 – 600
micrometers in diameter. The carbon granules are initially grinded in a disk mill until they come out as
fine, sand-like particles.

3.3.3 Sieving Particles


The carbon particles must be separated by size when grinded. A sieve is used to separate the
carbon particles and classify them by size. Sieves are stacked on top of each other with the largest pore
size sieve on top and the smallest on the bottom. For example, the 425 – 600 microns sieve is on top, 212
– 425 microns and then 125 – 212 microns. Anything below 100 micrometers in particle size is thrown
out and the rest is kept. The size of the particles used in the experiments are in the range of 212 – 425
micrometers.

49
Chapter 3: Experimental Apparatus

3.4 Microwave Unit Preparation


3.4.1 Cleaning the Microwave Cavity
The microwave should be clean and absent of any carbon particulates or debris inside the cavity.
If even a few particles are inside, the particles will heat up and melt the immediate area of the particles.
To clean the cavity, it was extensively wiped down with sanitary wipes.

3.4.2 Preparing and Installing Reaction Vessel


The quartz reaction vessel is reusable and is cleaned after each use. Three fritted discs are placed
inside the tube. This is the base and will hold the carbon particles upright. Anything less will cause the
disc level to be uneven will result in the carbon to fall to the bottom. The prepared reaction vessel with
the fritted discs and carbon should look like Figure 30 below.

Figure 30. Quartz vessel filled with carbon particles sitting on fritted discs

To install the prepared reaction vessel, it can be inserted from the bottom or the top of the
microwave chokes (assuming the microwave is in the vertical position). The muffle is placed in the cavity
in a way the holes on the muffle align with the holes on the chokes so that the quartz slides through the
chokes and the muffle.

Please see Appendix B for a detailed experimental setup procedure.

50
Chapter 3: Experimental Apparatus

3.4.3 Installing and Sealing Plumbing


Once the reaction vessel is placed in the chokes and muffle, the sealing connections can be
attached. The sealing equipment is that of in Figure 31 below where it will attach on to the flow adapter
connection. When tightened, the O-ring will compress onto the quartz tube where it will seal and prevent
any gas from escaping from both sides. Periodically, anti-seize paste is applied on to the threads in
between these connections. The anti-seize is used to prevent galling, seizing and corrosion, as well as
lubricating to ease assembly.

Figure 31. Ultra torr sealing configuration for bottom (left) and top (right) choke

Please see Appendix B for a detailed experimental setup procedure.

3.4.4 Flow Controller Calibration


Before any set of experiments, the flow controllers are calibrated with the gas it is assigned to be
used for so that accurate input flow rates can be utilized. Detailed steps for flow calibration are shown in
Appendix F.

51
Chapter 3: Experimental Apparatus

3.4.5 Filtration
A paper filter is placed in the exhaust to filter out any dust or particulates that manages to escape
the system. A new paper filter should be in place after each set of experiments. A secondary 2-micron
particulate filter is placed after the paper filter and right before the GC-TCD. Periodically, pressurized air
is flowed in the opposite direction of the intended flow direction of the particulate filter to blow out any
debris that may be clogging it. The images of these filters are shown in Figure 32 below.

Figure 32. Filter paper (left) and 2-micron particulate filter (right) are used to filter the exhaust

52
Chapter 3: Experimental Apparatus

3.5 Exhaust Analysis


3.5.1 GC-TCD
All gaseous products in the exhaust were directed to a gas chromatographer (GC) which is
equipped with a thermal conductivity detector (TCD) purchased from PerkinElmer. This reveals the
composition of the exhaust gas that’s been directed into the GC. The GC-TCD takes about 11 minutes to
analyze a sample, therefore the exhaust can only be analyzed every 11 minutes throughout the entire
experiment. The GC-TCD spectrum reveals the volume fraction of hydrogen, ethane, ethylene, acetylene,
nitrogen, oxygen, methane, carbon monoxide, carbon dioxide and many other species in the exhaust. The
composition of the exhaust gas sample is displayed on the Total Chrom Navigator. The operation
instruction for the GC-TCD is detailed in Appendix D.

Figure 33. GC-TCD

53
Chapter 3: Experimental Apparatus

3.5.2 Total Chrom Navigator


The total chrom navigator displays the results of the exhaust every 11 minutes. A sample result
has been shown in Figure 34. In this sample, hydrogen, nitrogen and methane peaks can be seen. The
higher the peak, the higher the concentration of the associated species. The concentration of the species in
a sample was obtained by comparing a peak value in a sample to the peak value of a species in a
calibration gas with a known concentration. For example, the concentration of hydrogen in Figure 34 can
be obtained by comparing its peak value to the peak value of hydrogen from a calibration gas of a known
hydrogen concentration. The unlabeled peak in between the hydrogen and nitrogen peak is due to a
disturbance caused by the components moving around inside the GC. This is normal to observe and
occurs in all results ever taken by the GC and can be ignored.

Figure 34. Exhaust results on Total Chrom Navigator showing hydrogen, nitrogen and methane peaks

54
Chapter 4: Experimental Results

Chapter 4: Experimental Results


Due to the strong bond between carbon and hydrogen, uncatalyzed methane pyrolysis requires a
very high temperature to break their bonds [108]. Low-temperature methane pyrolysis is feasible only in
the presence of catalysts [97]. However, the catalytic activity deactivates when the deposited carbon
covers their surface during pyrolysis. Despite the efforts taken for catalyst regeneration [97], catalyst
deactivation is still the most serious drawback of this technology regardless of catalyst type, carbonaceous
or molten metal [73], [109]. Although they are capable of continuous operation which produces hydrogen
and separable solid carbon as a by-product [61], the carbon is contaminated by molten reactor compounds
[62], limiting its potential applications. Scale-up challenges are another disadvantage of this technology
[110].

A new approach to continuously produce CO2-free hydrogen from catalyst-free methane


pyrolysis, exhibiting high methane conversion and excellent hydrogen selectivity is introduced. In this
method, methane pyrolysis is driven by microwave heating of carbon particles fluidized in a quartz
reactor (see Figure 35). Microwave heating is efficient and contactless [8] so that the heat can be
generated where it is required [9]. This pyrolysis system (see Figure 35A) is modular and can be deployed
wherever natural gas and electricity are available, avoiding the hydrogen distribution challenges
associated with its storage and transportation [11].

This chapter explores the results of experimental methane pyrolysis in a fluidized bed reactor
filled with carbon particles heated by microwaves (see Figure 35B). The fundamental procedure and
findings of this method is that the carbon particles absorb microwave energy, the gas temperature rises
due to particle-to-gas heat transfer providing the required energy for the pyrolysis reaction (see Figure
35C). The fluidization circulates the carbon particles which further enhances the heat transfer between the
particles and the gas [10]. In addition, the fluidized bed is a solution to blockage issues due to carbon
deposition in packed bed reactors [56]. Under these conditions, in roughly 500 cumulative operation
hours, methane decomposes into solid carbon and hydrogen exhibiting high methane conversion and

55
Chapter 4: Experimental Results

hydrogen selectivity [111]. Hydrogen leaves the reactor, and carbon deposits onto the existing carbon
particles inside the bed, resulting in particle growth (see Figure 35C).

Figure 35. Hydrogen production by microwave heating of a fluidized activated carbon particle bed. (A) Vertical microwave with
fluidized bed inside. (B) Fluidized bed reactor where methane flows from bottom to top. (C) Carbon particles absorb microwave
radiation and heat up for the methane pyrolysis reaction. At temperatures above 950℃, methane starts decomposing to hydrogen
and solid carbon. The initial particles grow in size due to carbon deposition onto their surface.

The experimental process involved pre-heating the carbon particles with 0.2 L/min of pure
nitrogen up to the desired temperature (around 1000°C). Once temperature stability was achieved, the
nitrogen flow rate was lowered to 0.1 L/min and methane was introduced at 0.1 L/min. A total flow rate
of 0.2 L/min is used throughout all experiments described below. After about 5 – 10 minutes of methane-
nitrogen flow, the GC-TCD takes a sample to measure. The GC-TCD takes a sample every 11 minutes so
data points can only be collected every 11 minutes. Based on the data collected from GC-TCD, the
methane-to-hydrogen conversion and hydrogen selectivity are calculated using equation (26) & (27)
respectively:

56
Chapter 4: Experimental Results

Methane conversion (%)

100 × (𝐶𝐻4 𝑚𝑜𝑙𝑎𝑟 𝑓𝑙𝑜𝑤 𝑟𝑎𝑡𝑒 (𝑖𝑛𝑙𝑒𝑡) − 𝐶𝐻4 𝑚𝑜𝑙𝑎𝑟 𝑓𝑙𝑜𝑤 𝑟𝑎𝑡𝑒 (𝑜𝑢𝑡𝑙𝑒𝑡))
= (26)
𝐶𝐻4 𝑚𝑜𝑙𝑎𝑟 𝑓𝑙𝑜𝑤 𝑟𝑎𝑡𝑒 (𝑖𝑛𝑙𝑒𝑡)

Hydrogen selectivity (%)

100 × 𝐻 𝑚𝑜𝑙𝑎𝑟 𝑓𝑙𝑜𝑤 𝑟𝑎𝑡𝑒 𝑖𝑛 𝐻2 (𝑜𝑢𝑡𝑙𝑒𝑡)


= (27)
𝐻 𝑚𝑜𝑙𝑎𝑟 𝑓𝑙𝑜𝑤 𝑟𝑎𝑡𝑒 𝑖𝑛 𝐶𝐻4 (𝑖𝑛𝑙𝑒𝑡) − 𝐻 𝑚𝑜𝑙𝑎𝑟 𝑓𝑙𝑜𝑤 𝑟𝑎𝑡𝑒 𝑖𝑛 𝐶𝐻4 (𝑜𝑢𝑡𝑙𝑒𝑡)

The following sections investigates the results of carbon deactivation, particle growth, methane to
hydrogen conversion rates, methane decomposition species, carbon mass balance and effect of dilution all
in more detail.

4.1 Carbon Particle Heating, Deactivation and Growth


Carbon-containing material is well known to be a good microwave absorber [102]. The pyrolysis
process produces a continuous supply of carbon material for microwave absorption. However, for this
experiment, initial seed particles are needed to begin the process. Activated carbon of size, 212 – 425
microns, is used as the initial seed particles as it is a suitable microwave receptor [103]. Approximately
45 grams of activated carbon particles are pre-heated to a temperature of about 917°C. Once the
temperature stabilized, methane was introduced with methane conversion starting almost immediately.
Figure 36 displays that at an average temperature of 917°C, the methane conversion drops rapidly from
near 100% to less than 10% after about two hours and stabilizes after three hours. Figure 36 indicates that
if the experiment were to continue, the conversion rate may drop even more. This was investigated by
using the same carbon particles for a total of roughly 500 hours of pyrolysis at various methane
conversion rates resulting in a 5.3% conversion of methane at 920°C. This process deactivates the
catalytic effect of the carbon particles due to carbon deposition from the methane where the conversion
rate becomes steady after about 3 – 4 hours. Scanning transmission electron microscope (SEM) images of
the carbon particles before and after pyrolysis were taken (see Figure 37). The SEM images indicate that
before pyrolysis, activated carbon is highly porous and fluffy (see Figure 37A). In addition, Figure 37
also shows the increase in size before and after pyrolysis from the original activated carbon to the covered
carbon. The high surface area and significant active sites in activated carbon accelerates methane
pyrolysis at temperatures below 900℃ [100]. However, the carbon deposition from methane pyrolysis
fills the pores and covers the surface and active sites (see Figure 37B), leading to significant surface area
reduction. Please refer to Appendix G for additional SEM images on activated carbon particles before and
after methane decomposition. Brunauer-Emmett-Teller (BET) surface area analysis reveals that the

57
Chapter 4: Experimental Results

covered carbon has a specific surface area of 0.4 m2/g which is three orders of magnitude less than the
specific surface area of activated carbon of 846 m2/g. The bulk density of the carbon before and after
pyrolysis increased from 500 kg/m3 to 865 kg/m3. This reduction of surface area causes the methane
conversion rate to drop rapidly [69], [95], [100] from near 100% to less than 10% in about 90 minutes.
This indicates that when the conversion rate stabilizes, the surface area of the covered particles is
minimized. As a result, the effect of carbon particle deactivation means that gas temperature becomes the
dominant parameter that controls the conversion rate of methane, transitioning it from catalytic methane
decomposition to thermal methane decomposition [53].

It should be emphasized that if there’s insufficient fluidization or if it’s a packed bed, carbon
deposition will bridge the gaps in between the carbon particles causing it to agglomerate and produce a
much larger particle of a few centimeters in size (see Figure 38). Fluidization is heavily dependent on the
particle size. If the particles of a few hundred microns in size clump together to form a size that’s a few
centimeters, the uniform flow will be heavily disrupted, and the size would be far too large to fluidize
effectively. Figure 38 shows that due to insufficient fluidization in certain regions of the reactor, the
carbon particles agglomerated to form a much larger particle. To prevent this, higher flow rates should be
used as this will increase the space in between the particles and allow more movement in the reactor.

Figure 36. Effect of carbon deposition on the surface of activated carbon expressed by time versus methane conversion rate.
Methane conversion rate starts at ~100%, but quickly declines reaching a steady state after almost three hours of run time.

58
Chapter 4: Experimental Results

Figure 37. Scanning electron microscope (SEM) images of activated carbon particles (212 - 425 microns) before pyrolysis (A)
and after pyrolysis (B) from 1mm, 300 micrometer and 50 micrometer magnification respectively

59
Chapter 4: Experimental Results

Carbon agglomerate formation after methane pyrolysis

Figure 38. Carbon particle agglomerates from carbon deposition bridging particles due to poor fluidization

The covered carbon particles are used as the microwave receptor instead of periodically changing
it with fresh activated carbon. The gas temperature is a strong function of the applied microwave energy
and the microwave absorption capacity of carbon particles. Therefore, it’s necessary to investigate the
microwave absorption capabilities of the covered carbon. Figure 39 and Figure 40 displays the
microwave-heating of covered particles indicating that they still possess high microwave absorption
capabilities by reaching ~1200°C after 20 min. The temperature of the carbon bed is controlled by the
power output dial on the microwave. The following results have all been obtained using covered carbon.

Figure 39. Gas temperature in contact with covered particles absorbing microwave energy

60
Chapter 4: Experimental Results

Figure 40. Heated quartz reactor

61
Chapter 4: Experimental Results

4.2 Methane Pyrolysis Conversion Rates


To evaluate the effect of temperature on methane conversion, A 50-50% methane nitrogen
mixture with a flow rate of 0.2 L/min is introduced into a 22 mm ID reactor filled with deactivated
(covered) carbon particles heated in a temperature range of 750℃ to 1216℃ (please note that covered
carbon and deactivated carbon will be used interchangeably as it refers to the same thing). Deactivated
carbon particles are used to minimize any catalytic effect so that the methane conversion rate would only
be affected by the temperature. The temperature would gradually increase by controlling the microwave
power and the exhaust was measured every 11 minutes by the GC-TCD. The methane conversion rate and
hydrogen selectivity were calculated using equation (26) and (27) respectively. The results of this are
displayed in Figure 41 where the symbols and solid line respectively show the experimental and
modelling results. As can be seen from the results, methane conversion starts at just under 900°C,
increasing rapidly at ~950℃ and reaches 90% at 1216℃. This indicates that temperature is playing a
strong role on the decomposition of methane. Hydrogen selectivity is modest at around temperatures of
900 – 1000°C. At temperatures of over 1000°C, more than 90% hydrogen selectivity was achieved. At
temperatures of at least 1100°C, almost 100% hydrogen selectivity is achieved. A 90% hydrogen
selectivity means that most of the methane conversion resulted in an output of pure hydrogen gas with the
rest being other species containing hydrogen atoms (major and minor species of methane conversion is
discussed in the next section). Higher methane conversion and hydrogen purity can be achieved with
increased temperatures. These results indicate that the deactivated carbon can be reused as the microwave
receptor without needing the original activated carbon with the caveat of requiring higher temperatures.
The modeling results (solid line) follow the experimental trend in methane pyrolysis and hydrogen
selectivity but due to the assumptions and simplifications made in the model, it underpredicts the
conversion rate.

62
Chapter 4: Experimental Results

Figure 41. Methane conversion rates and hydrogen selectivity over a range of temperatures, pressure of 1 atm and estimated
residence time of ~5 seconds

The reactor model includes detailed chemical kinetics of methane pyrolysis coupled with a
sectional population balance model. To this end, a plug flow code (NanoPFR) [112] is modified to
simulate the cogeneration of hydrogen and carbon particles in a plug flow reactor based on a one-
dimensional geometry assumption (see Appendix H). NanoPFR code solves a mass conservation equation
for each species, energy, and momentum equations over small intervals. The ideal gas law is utilized to
relate the pressure to the density in the aforementioned governing equations. The microwave heating
energy is modeled via a fixed temperature based on the reactor temperature provided by the experimental
measurements. The detailed CALTECH kinetic mechanism file [113] is used for the gas-phase chemistry.
There are 192 species and 1116 reactions in the mechanism. A fixed pivot approach is employed for the
sectional population balance model [114], [115], the Smoluchowski equation [116], to simulate carbon
particle inception and surface growth. The inception model and the surface growth are further discussed
in Appendix H. The solid lines in Figure 41 represent the numerical results, indicating a reasonable
agreement with experimental work.

63
Chapter 4: Experimental Results

4.3 Major and Minor Species


Gaseous products are classified into major and minor species. The major species in the exhaust
consists of hydrogen, unreacted methane and nitrogen. Depending on experimental conditions, ~98% to
>99.9% of products are hydrogen, unreacted methane and nitrogen (see Figure 42 (left)). The remainder
(less than 2% by mole) are minor species consisting of ethane, ethylene, and acetylene. The minor
hydrocarbons start forming at around 820℃ with a maximum mole fraction at just over 1000℃ and
diminishing after 1200℃ as can be seen in Figure 42 (right). This is beneficial if large amounts of pure
hydrogen is to be produced as the minor species diminish after 1200°C. The proposed mechanism for
minor species formation suggests that ethane forms first because of methyl radicals combining, and the
ethylene and acetylene form through ethane and ethylene decomposition, respectively [117]. Acetylene is
consumed by addition to the surface of the particles through the hydrogen abstraction carbon addition
(HACA) mechanism [118] or the formation of polycyclic aromatic hydrocarbons (PAHs) [119]. For
major species, the experimental results (symbols) follow the trends of the modelling results (solid, dash
and dotted line). For minor species, the model underpredicts the experimental results which may be due to
the simplifications made in the model such as assuming plug flow and a fixed temperature.

Figure 42. Major species (left) and minor species (right) from methane pyrolysis over a range of temperatures. The symbols
represent the experimental results and the solid, dash and dotted lines represent the modeling results

64
Chapter 4: Experimental Results

4.4 Carbon Mass Balance


The products of methane pyrolysis can be divided into two groups: solid carbon and gases. In this
experiment, a 100% methane and 50-50 methane-nitrogen mix were introduced to deactivated carbon
particles to study where all the carbon goes. To track the carbon during methane pyrolysis, the mass of
carbon particles and a paper filter in the exhaust were measured before and after experiments. The paper
filter is used to track and measure the amount of carbon nanoparticle deposition. The solid carbon was
measured under two average temperatures, 947℃ and 1045℃. Table 4 reports the comparison between
measured solid carbon and the carbon mass calculated based on the gas species in the exhaust. The results
show that there is a trivial amount (~14 mg on average) of carbon nanoparticles depositing on the filter,
and almost all the solid carbon (>99%) has formed on the granules. The measured values are similar to
the theoretical amount of carbon produced indicating that most of the carbon has been accounted for.
Therefore, it is fair to assume the difference between the weight of the pellets before and after each
experiment would be equal to the total produced solid carbon.

Table 4. Carbon mass balance experiments under two different inlet mixture and temperature conditions

Inlet mixture 100% CH4 50% CH4 + 50% N2

Average reactor temperature (°C) 947 1045

Measured mass of carbon deposited on 0.9 2.7


existing particles (g)

Calculated mass of deposited carbon using 0.899 2.695


GC-TCD data (g)

Error (%) 0.11 0.19

Carbon on the filter (g) ~0.008 ~0.020

65
Chapter 4: Experimental Results

4.5 Thin-Film Carbon


Aside from the growth of the carbon granules due to carbon deposition from the pyrolyzed
methane, another type of carbon forms. On the inner walls of the reactor, a thin sheet of flaky reflective
carbon, similar to aluminum foil, deposits on the wall as shown in Figure 43 and Figure 44. This type of
carbon is reflective to light and microwaves. The formation of this prevents long term use of the
microwave as the temperature starts decreasing after a few hours of operation at 50% methane. This film
is also electrically conductive which isn’t the case with the carbon depositing on the carbon granules.
SEM images of the carbon film is shown in Figure 45 where it can be seen that it’s quite flaky with
jagged edges.

This thin film carbon does appear to be a phenomenon in other methane pyrolysis experiments
[80], [120]. Based on Parker’s work [121], this thin film carbon could potentially be pyrolytic carbon as
pyrolytic carbon film was mentioned to form in the temperature regions of 900°C – 1300°C, while also
being optically reflective with a thickness of about 200 nm thick [121]. When the thin film carbon is
attempted to be extracted, it completely breaks down as soon as it’s disturbed. If a simpler extraction and
controlled formation method is formulated, this type of carbon could see applications as an electrically
conductive film or paste in electronic components.

Figure 43. Thin film carbon formation in reactor

66
Chapter 4: Experimental Results

Figure 44. Collected thin film carbon

Figure 45. SEM image of thin film carbon at 50 micrometers

67
Chapter 4: Experimental Results

4.6 Methane Dilution


The methane was diluted with nitrogen for the safety of the experiment in the laboratory. This
raises an interesting question on the effect dilution may have on the methane conversion rate and allow a
better understanding on what would happen if natural gas were to be used since natural gas isn’t pure
methane. The effect of nitrogen addition was investigated by diluting methane to 10% methane and 90%
nitrogen and increasing the methane concentration until 100%. This was done in temperature ranges of
1000°C to 1160°C using deactivated carbon particles. The results of this are reported in Figure 46 below
which show that the methane conversion rates of the diluted methane are similar to that of Figure 41.
These experiments ultimately indicate that methane dilution with nitrogen has a negligible effect on the
methane conversion rate. This can be good news when trying to implement natural gas in the process
rather than pure methane. But since natural gas contains other compounds as well in addition to nitrogen,
it is unknown what affect the remaining compounds may have on the conversion rate. Deng has reported
an effect of nitrogen on methane conversion over activated carbon via microwave heating [122]. The
effect of nitrogen in Deng’s results are not obvious as it doesn’t show any conclusive patterns.
Regardless, the negligible effect of nitrogen shown in Figure 46 may be due to using deactivated carbon
compared to Deng’s activated carbon thus not making it a one-to-one comparison.

Figure 46. Effect of dilution on methane conversion rate at a pressure of 1 atm and estimated residence time of ~5 seconds

68
Chapter 4: Experimental Results

4.7 Activation Energy


The effective activation energy of the reaction was estimated as described in Appendix I. The
experimental activation energy of 224 kJ/mol (see Figure 47) is a little higher than the range of 160 to 201
kJ/mol for activated carbon catalysts at 600 – 900°C [93] but ~40% lower than uncatalyzed methane
pyrolysis, reported as 370 kJ/mol [123]. This comparison suggests that the carbon that covers the initial
activated carbon particles does get slightly deactivated, but it may have autocatalytic effects since the
activation energy did not get into the 300 range [99], [124]. This is very beneficial as there’s no need for
catalyst regeneration or catalyst replacement. Over long periods of methane decomposition, the carbon
particles will presumably continue to grow thus affecting fluidization. Periodically, the covered carbon
granules will need to be removed and ground back down to its original size range of 212 – 425 microns. If
this is done over a few iterations, none of the original seed particles will remain.

Figure 47. Activation energy of covered carbon particles at temperatures of 600 – 900°C, pressure of 1 atm and estimated
residence time of ~5 seconds

69
Chapter 5: Conclusion and Future Work

Chapter 5: Conclusion and Future Work


5.1 Conclusion
There is a lot of buzz around hydrogen due to its nature of clean combustion and has thus been
termed as the “fuel of the future” by its many proponents. A shift into a “hydrogen economy” where
hydrogen is used to decarbonize heavy duty industries such as steel, cement, plastic, long-haul transport,
shipping and aviation can only be done if hydrogen can be produced through clean methods. This thesis
investigates a novel approach to CO2-free hydrogen production, known as turquoise hydrogen. In this
method, microwave radiation is shone on carbon particles (of size 212 – 425 microns) that is fluidized by
incoming methane gas where it will heat up to temperatures of 1200°C and decompose into pure solid
carbon and hydrogen. Microwave absorption by the carbon particles which are in contact with methane,
provide the energy for the pyrolysis process. It has been found that methane starts decomposing at around
900°C with a sharp rise at 950℃ and achieving 90% methane to hydrogen conversion at 1216℃. The
hydrogen selectivity is achieved to be more than 90% at temperatures over 1000°C.

Solid carbon is a major by-product of this process with minor by-product species consisting of
ethane, ethylene, and acetylene composing of less than 2% of total products. All minor species completely
diminish at temperatures over 1200°C and convert into hydrogen. The mass balance under two different
conditions suggests that the solid carbon formed through the pyrolysis process deposits on the existing
particles and only a trivial amount of carbon nanoparticles is produced. More than 500 cumulative hours
of operation show that covering the existing particles with the formed carbon does not degrade their
microwave absorption property. Therefore, the carbon can be reused in the pyrolysis process. Since there
are no metal catalysts in this process, the formed carbon has a high purity and depending on its
morphology, it can be used in a variety of applications such as graphene formation, steelmaking, battery
cathodes, etc. The thin-film carbon prevents long experiment run-times as it’s microwave reflective. It’s
also highly electrically conductive and could see applications in electronic components as a film or paste.

70
Chapter 5: Conclusion and Future Work

The effective activation energy of 224 kJ/mol was calculated and lies a little higher than the range
for activated carbon catalysts but lower than the range for uncatalyzed methane pyrolysis, suggesting the
presence of autocatalytic effects for produced carbon. This autocatalytic effect disposes the need for
catalyst regeneration or catalyst replacement which is a major issue in this field. Periodically, due to
particle growth, the covered carbon granules will need to be removed and grind back down to its original
size range of 212 – 425 microns. The continual process of carbon growth and grinding over a few
iterations will leave none of the original seed particles and only the produced carbon will remain.

With more than 90% methane to hydrogen conversion and hydrogen selectivity, this method
could potentially be able to produce sizeable quantities of hydrogen with high purity and marketable
carbon, assuming the cost to produce hydrogen per kilogram is competitive with SMR. The compact
design of this system shows its potential to be used as a modular and scalable unit anywhere with access
to natural gas and electricity for localized hydrogen production, bypassing challenges associated with
hydrogen distribution, storage and outright avoiding CO2 capture and sequestration (CCS).

5.2 Future Work


5.2.1 Experimental Apparatus Improvements
Some improvements that can be made to the experimental apparatus are:

• Larger chokes to allow larger quartz tubes and more mass flow rates. Larger flow rates would
allow more room to study fluidization effects
• Larger microwave for more power
• Better fans for increases circulation inside plexiglass box and experiment room
• Continuous measurements of exhaust rather than measurements every 11 minutes in GC-TCD
• Sturdier reactor that can withstand higher temperatures
• Power consumption meter to establish the power used in the process

5.2.2 Future Research


Future work that is of interest is listed below:

• Control of thin film carbon formation


• Longer run time periods to study clogging
• Larger quartz ID for larger mass flow rates. Larger mass flow rates would allow more room to
study fluidization effects
• Residence time studies

71
Chapter 5: Conclusion and Future Work

• Energy balance to study the overall efficiency of the system (hydrogen out/energy in)
• Morphology study on the carbon. What kind of carbon is being produced?
• What applications can the produced carbon be used in (covered carbon and thin-film carbon)?
• Effect of pressure on conversion rate
• Discovery of mechanism for carbon formation
• Particle size distribution
• Apply better hardware cooling
• Natural gas pyrolysis vs pure methane pyrolysis
• Temperature distribution due to thermal effects of microwave
• Microwave vs furnace. How are the results different if the furnace uses the same conditions
• Effect of frequency and input microwave power
• Performance of fluidization
• Does even higher temperatures change carbon (or hydrogen) formation
• Techno-economic assessment
• Effect of reactor material

72
References
[1] J. M. Ogden, “Hydrogen: The Fuel of the Future?,” Phys Today, vol. 55, no. 4, p. 69, Jan. 2007, doi:
10.1063/1.1480785.

[2] G. W. Crabtree, M. S. Dresselhaus, and M. v. Buchanan, “The Hydrogen Economy,” Phys Today,
vol. 57, no. 12, p. 39, Jan. 2007, doi: 10.1063/1.1878333.

[3] Masson-Delmotte, V., P. Zhai, and H.-O. Portner, “Summary for Policymakers. In: Global Warming
of 1.5°C. An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial
levels and related global greenhouse gas emission pathways, in the context of strengthening the
global response to the threat of climate change, sustainable development,
and efforts to eradicate poverty,” Cambridge University Press, Jun. 2018. doi:
10.1017/9781009157940.001.

[4] J. Diab, L. Fulcheri, V. Hessel, V. Rohani, and M. Frenklach, “Why turquoise hydrogen will Be a
game changer for the energy transition,” Int J Hydrogen Energy, vol. 47, no. 61, pp. 25831–
25848, Jul. 2022, doi: 10.1016/J.IJHYDENE.2022.05.299.

[5] N. Z. Muradov and T. N. Veziroǧlu, “From hydrocarbon to hydrogen–carbon to hydrogen


economy,” Int J Hydrogen Energy, vol. 30, no. 3, pp. 225–237, Mar. 2005, doi:
10.1016/J.IJHYDENE.2004.03.033.

[6] G. Marbán and T. Valdés-Solís, “Towards the hydrogen economy?,” Int J Hydrogen Energy, vol.
32, no. 12, pp. 1625–1637, Aug. 2007, doi: 10.1016/J.IJHYDENE.2006.12.017.

[7] L. Weger, A. Abánades, and T. Butler, “Methane cracking as a bridge technology to the hydrogen
economy,” Int J Hydrogen Energy, vol. 42, no. 1, pp. 720–731, Jan. 2017, doi:
10.1016/J.IJHYDENE.2016.11.029.

[8] J. M. Serra et al., “Hydrogen production via microwave-induced water splitting at low
temperature,” Nature Energy 2020 5:11, vol. 5, no. 11, pp. 910–919, Nov. 2020, doi:
10.1038/s41560-020-00720-6.

[9] J. Sun, W. Wang, and Q. Yue, “Review on Microwave-Matter Interaction Fundamentals and
Efficient Microwave-Associated Heating Strategies,” Materials (Basel), vol. 9, no. 4, 2016, doi:
10.3390/MA9040231.

[10] D. Kunii and Octave. Levenspiel, Fluidization Engineering, Second Edition. Stoneham:
Butterworth-Heinemann, 1991.

[11] A. M. Abdalla, S. Hossain, O. B. Nisfindy, A. T. Azad, M. Dawood, and A. K. Azad, “Hydrogen


production, storage, transportation and key challenges with applications: A review,” Energy
Convers Manag, vol. 165, pp. 602–627, Jun. 2018, doi: 10.1016/J.ENCONMAN.2018.03.088.

[12] “NIST Chemistry WebBook.” https://webbook.nist.gov/chemistry/ (accessed Jul. 25, 2022).

[13] V. Madadi Avargani, S. Zendehboudi, N. M. Cata Saady, and M. B. Dusseault, “A comprehensive


review on hydrogen production and utilization in North America: Prospects and challenges,”
Energy Convers Manag, vol. 269, p. 115927, Oct. 2022, doi: 10.1016/J.ENCONMAN.2022.115927.

73
[14] International Energy Agency, “The Future of Hydrogen - Seizing today’s opportunities,” 2019.

[15] International Energy Agency, “Hydrogen – Analysis,” 2021.


https://www.iea.org/reports/hydrogen (accessed Aug. 01, 2022).

[16] A. M. Aitani, “Oil Refining and Products,” Encyclopedia of Energy, pp. 715–729, Jan. 2004, doi:
10.1016/B0-12-176480-X/00259-X.

[17] International Energy Agency, “Oil 2021 – Analysis - IEA,” 2021. https://www.iea.org/reports/oil-
2021 (accessed Aug. 28, 2022).

[18] McKinsey, “Global oil outlook to 2040,” 2021.

[19] Organization of the Petroleum Exporting Countries (OPEC), “World Oil Outlook 2045 Organization
of the Petroleum Exporting Countries,” 2021.

[20] M. Appl, “Ammonia,” Ullmann’s Encyclopedia of Industrial Chemistry, Dec. 2006, doi:
10.1002/14356007.A02_143.PUB2.

[21] R. de María, I. Díaz, M. Rodríguez, and A. Sáiz, “Industrial methanol from syngas: Kinetic study
and process simulation,” International Journal of Chemical Reactor Engineering, vol. 11, no. 1, pp.
469–477, Aug. 2013, doi: 10.1515/IJCRE-2013-0061/MACHINEREADABLECITATION/RIS.

[22] F. Patisson and O. Mirgaux, “Hydrogen ironmaking: How it works,” Metals (Basel), vol. 10, no. 7,
pp. 1–15, Jul. 2020, doi: 10.3390/met10070922.

[23] X. A. Huang, K. W. Ng, L. Giroux, and M. Duchesne, “Carbonaceous Material Properties and Their
Interactions with Slag During Electric Arc Furnace Steelmaking,” Metallurgical and Materials
Transactions B: Process Metallurgy and Materials Processing Science, vol. 50, no. 3, pp. 1387–
1398, Jun. 2019, doi: 10.1007/s11663-019-01569-1.

[24] Hydrogen Council, “Hydrogen scaling up a sustainable pathway for the global energy transition,”
2017, Accessed: Sep. 04, 2022. [Online]. Available: www.hydrogencouncil.com.

[25] M. Hermesmann and T. E. Müller, “Green, Turquoise, Blue, or Grey? Environmentally friendly
Hydrogen Production in Transforming Energy Systems,” Prog Energy Combust Sci, vol. 90, p.
100996, May 2022, doi: 10.1016/J.PECS.2022.100996.

[26] IEAGHG, “Techno - Economic Evaluation of SMR Based Standalone (Merchant) Hydrogen Plant
with CCS,” Feb. 2017. Accessed: Sep. 20, 2022. [Online]. Available: www.ieaghg.org

[27] “Commercial-scale methane pyrolysis: The key to carbon-negative hydrogen - Energy Monitor,”
Apr. 04, 2022. https://www.energymonitor.ai/sponsored/commercial-scale-methane-pyrolysis-
the-key-to-carbon-negative-hydrogen (accessed Nov. 17, 2022).

[28] R. Carapellucci and L. Giordano, “Steam, dry and autothermal methane reforming for hydrogen
production: A thermodynamic equilibrium analysis,” J Power Sources, vol. 469, p. 228391, Sep.
2020, doi: 10.1016/J.JPOWSOUR.2020.228391.

74
[29] M. Steinberg and H. C. Cheng, “Modern and prospective technologies for hydrogen production
from fossil fuels,” Int J Hydrogen Energy, vol. 14, no. 11, pp. 797–820, 1989, doi:
https://doi.org/10.1016/0360-3199(89)90018-9.

[30] Enbridge Gas, “Learn About Natural Gas | Enbridge Gas.” https://www.enbridgegas.com/about-
enbridge-gas/learn-about-natural-gas (accessed Sep. 11, 2022).

[31] C. H. Bartholomew, “Mechanisms of Nickel Catalyst Poisoning,” Stud Surf Sci Catal, vol. 34, no. C,
pp. 81–104, Jan. 1987, doi: 10.1016/S0167-2991(09)60352-9.

[32] J. N. Armor, “The multiple roles for catalysis in the production of H2,” Appl Catal A Gen, vol. 176,
no. 2, pp. 159–176, Jan. 1999, doi: 10.1016/S0926-860X(98)00244-0.

[33] J. R. Rostrup-Nielsen, “Activity of nickel catalysts for steam reforming of hydrocarbons,” J Catal,
vol. 31, no. 2, pp. 173–199, Nov. 1973, doi: 10.1016/0021-9517(73)90326-6.

[34] C. J. Liu, J. Ye, J. Jiang, and Y. Pan, “Progresses in the Preparation of Coke Resistant Ni-based
Catalyst for Steam and CO2 Reforming of Methane,” ChemCatChem, vol. 3, no. 3, pp. 529–541,
Mar. 2011, doi: 10.1002/CCTC.201000358.

[35] R. M. Cuéllar-Franca and A. Azapagic, “Carbon capture, storage and utilisation technologies: A
critical analysis and comparison of their life cycle environmental impacts,” Journal of CO2
Utilization, vol. 9, pp. 82–102, Mar. 2015, doi: 10.1016/J.JCOU.2014.12.001.

[36] International Energy Agency and United Nations Industrial Development Organization,
“Technology Roadmap Carbon Capture and Storage in Industrial Applications,” 2011.

[37] B. Singh, A. H. Strømman, and E. G. Hertwich, “Comparative life cycle environmental assessment
of CCS technologies,” International Journal of Greenhouse Gas Control, vol. 5, no. 4, pp. 911–921,
Jul. 2011, doi: 10.1016/J.IJGGC.2011.03.012.

[38] A. Bello and R. O. Idem, “Pathways for the Formation of Products of the Oxidative Degradation of
CO 2-Loaded Concentrated Aqueous Monoethanolamine Solutions during CO 2 Absorption from
Flue Gases,” 2005, doi: 10.1021/ie049329.

[39] Working Group IIII of the Intergovernmental Panel on Climate Change, “IPCC Special Report on
Carbon Dioxide Capture and Storage,” New York, NY, USA, 2005.

[40] Cal Cooper and CO2 Capture Project, “A Technical Basis For Carbon Dioxide Storage,” 2009.

[41] K. Brown, S. Whittaker, M. Wilson, W. Srisang, H. Smithson, and P. Tontiwachwuthikul, “The


history and development of the IEA GHG Weyburn-Midale CO2 Monitoring and Storage Project in
Saskatchewan, Canada (the world largest CO2 for EOR and CCS program),” Petroleum, vol. 3, no.
1, pp. 3–9, Mar. 2017, doi: 10.1016/J.PETLM.2016.12.002.

[42] R. W. Howarth and M. Z. Jacobson, “How green is blue hydrogen?,” Energy Sci Eng, vol. 9, no. 10,
pp. 1676–1687, Oct. 2021, doi: 10.1002/ESE3.956.

[43] S. Trasatti, “Water electrolysis: who first?,” Journal of Electroanalytical Chemistry, vol. 476, no. 1,
pp. 90–91, Oct. 1999, doi: 10.1016/S0022-0728(99)00364-2.

75
[44] P. Millet, “Fundamentals of water electrolysis,” Electrochemical Power Sources: Fundamentals,
Systems, and Applications, pp. 37–62, 2022, doi: 10.1016/B978-0-12-819424-9.00002-1.

[45] R. G. Ehl, A. J. Ihde, J. Ns, and J. Berzelius, “Faraday’s electrochemical laws and the determination
of equivalent weights”, Accessed: Oct. 02, 2022. [Online]. Available:
https://pubs.acs.org/sharingguidelines

[46] A. Ursúa, L. M. Gandía, and P. Sanchis, “Hydrogen production from water electrolysis: Current
status and future trends,” in Proceedings of the IEEE, 2012, vol. 100, no. 2, pp. 410–426. doi:
10.1109/JPROC.2011.2156750.

[47] S. Sebbahi, N. Nabil, A. Alaoui-Belghiti, S. Laasri, S. Rachidi, and A. Hajjaji, “Assessment of the
three most developed water electrolysis technologies: Alkaline Water Electrolysis, Proton
Exchange Membrane and Solid-Oxide Electrolysis,” Mater Today Proc, vol. 66, pp. 140–145, Jan.
2022, doi: 10.1016/J.MATPR.2022.04.264.

[48] S. A. Grigoriev, V. N. Fateev, D. G. Bessarabov, and P. Millet, “Current status, research trends, and
challenges in water electrolysis science and technology,” Int J Hydrogen Energy, vol. 45, no. 49,
pp. 26036–26058, Oct. 2020, doi: 10.1016/J.IJHYDENE.2020.03.109.

[49] J. Gonzalez-Aguilar, I. Dème, L. Fulcheri, G. Flamant, T. M. Gruenberger, and B. Ravary,


“Comparison of Simple Particle-Radiation Coupling Models Applied on a Plasma Black Process,”
Plasma Chemistry and Plasma Processing 2004 24:4, vol. 24, no. 4, pp. 603–623, Dec. 2004, doi:
10.1007/S11090-004-7935-5.

[50] N. Sánchez-Bastardo, R. Schlögl, and H. Ruland, “Methane Pyrolysis for CO2-Free H2 Production:
A Green Process to Overcome Renewable Energies Unsteadiness,” Chemie Ingenieur Technik, vol.
92, no. 10, pp. 1596–1609, Oct. 2020, doi: 10.1002/CITE.202000029.

[51] T. Keipi, K. E. S. Tolvanen, H. Tolvanen, and J. Konttinen, “Thermo-catalytic decomposition of


methane: The effect of reaction parameters on process design and the utilization possibilities of
the produced carbon,” Energy Convers Manag, vol. 126, pp. 923–934, Oct. 2016, doi:
10.1016/J.ENCONMAN.2016.08.060.

[52] J. R. Fincke, R. P. Anderson, T. A. Hyde, and B. A. Detering, “Plasma pyrolysis of methane to


hydrogen and carbon black,” Ind Eng Chem Res, vol. 41, no. 6, pp. 1425–1435, 2002, doi:
10.1021/ie010722e.

[53] S. Schneider, S. Bajohr, F. Graf, and T. Kolb, “State of the Art of Hydrogen Production via Pyrolysis
of Natural Gas,” ChemBioEng Reviews, vol. 7, no. 5. Wiley-Blackwell, pp. 150–158, Oct. 01, 2020.
doi: 10.1002/cben.202000014.

[54] G. Chen, X. Tu, G. Homm, and A. Weidenkaff, “Plasma pyrolysis for a sustainable hydrogen
economy,” Nat Rev Mater, Apr. 2022, doi: 10.1038/S41578-022-00439-8.

[55] A. M. Amin, E. Croiset, and W. Epling, “Review of methane catalytic cracking for hydrogen
production,” Int J Hydrogen Energy, vol. 36, no. 4, pp. 2904–2935, Feb. 2011, doi:
10.1016/j.ijhydene.2010.11.035.

76
[56] A. Abánades et al., “Experimental analysis of direct thermal methane cracking,” Int J Hydrogen
Energy, vol. 36, no. 20, pp. 12877–12886, Oct. 2011, doi: 10.1016/J.IJHYDENE.2011.07.081.

[57] M. Plevan et al., “Thermal cracking of methane in a liquid metal bubble column reactor:
Experiments and kinetic analysis,” Int J Hydrogen Energy, vol. 40, no. 25, pp. 8020–8033, Jul.
2015, doi: 10.1016/J.IJHYDENE.2015.04.062.

[58] T. Geißler et al., “Hydrogen production via methane pyrolysis in a liquid metal bubble column
reactor with a packed bed,” Chemical Engineering Journal, vol. 299, pp. 192–200, Sep. 2016, doi:
10.1016/J.CEJ.2016.04.066.

[59] M. Serban, M. A. Lewis, C. L. Marshall, and R. D. Doctor, “Hydrogen Production by Direct Contact
Pyrolysis of Natural Gas,” 2003, doi: 10.1021/ef020271q.

[60] B. Parkinson, J. W. Matthews, T. B. McConnaughy, D. C. Upham, and E. W. McFarland, “Techno-


Economic Analysis of Methane Pyrolysis in Molten Metals: Decarbonizing Natural Gas,” Chem Eng
Technol, vol. 40, no. 6, pp. 1022–1030, Jun. 2017, doi: 10.1002/CEAT.201600414.

[61] D. C. Upham et al., “Catalytic molten metals for the direct conversion of methane to hydrogen
and separable carbon,” Science (1979), vol. 358, no. 6365, pp. 917–921, Nov. 2017, doi:
10.1126/science.aao5023.

[62] N. Rahimi et al., “Solid carbon production and recovery from high temperature methane pyrolysis
in bubble columns containing molten metals and molten salts,” Carbon N Y, vol. 151, pp. 181–
191, Oct. 2019, doi: 10.1016/j.carbon.2019.05.041.

[63] T. Geißler et al., “Experimental investigation and thermo-chemical modeling of methane pyrolysis
in a liquid metal bubble column reactor with a packed bed,” Int J Hydrogen Energy, vol. 40, no.
41, pp. 14134–14146, Nov. 2015, doi: 10.1016/J.IJHYDENE.2015.08.102.

[64] C. Palmer et al., “Methane Pyrolysis with a Molten Cu-Bi Alloy Catalyst,” ACS Catal, vol. 9, no. 9,
pp. 8337–8345, Sep. 2019, doi:
10.1021/ACSCATAL.9B01833/ASSET/IMAGES/LARGE/CS9B01833_0009.JPEG.

[65] S. Rodat, S. Abanades, J. L. Sans, and G. Flamant, “A pilot-scale solar reactor for the production of
hydrogen and carbon black from methane splitting,” Int J Hydrogen Energy, vol. 35, no. 15, pp.
7748–7758, Aug. 2010, doi: 10.1016/J.IJHYDENE.2010.05.057.

[66] N. Muradov, “Thermocatalytic CO2-free production of hydrogen from hydrocarbon fuels,”


Proceedings of the 2000 US DOE Hydrogen Program Review, 2000.

[67] M. H. Kim et al., “Hydrogen production by catalytic decomposition of methane over activated
carbons: kinetic study,” Int J Hydrogen Energy, vol. 29, no. 2, pp. 187–193, Feb. 2004, doi:
10.1016/S0360-3199(03)00111-3.

[68] K. K. Lee, G. Y. Han, K. J. Yoon, and B. K. Lee, “Thermocatalytic hydrogen production from the
methane in a fluidized bed with activated carbon catalyst,” Catal Today, vol. 93–95, pp. 81–86,
Sep. 2004, doi: 10.1016/J.CATTOD.2004.06.080.

77
[69] A. M. Dunker, S. Kumar, and P. A. Mulawa, “Production of hydrogen by thermal decomposition of
methane in a fluidized-bed reactor—Effects of catalyst, temperature, and residence time,” Int J
Hydrogen Energy, vol. 31, no. 4, pp. 473–484, Mar. 2006, doi: 10.1016/J.IJHYDENE.2005.04.023.

[70] J. Zeng, M. Tarazkar, T. Pennebaker, M. J. Gordon, H. Metiu, and E. W. McFarland, “Catalytic


Methane Pyrolysis with Liquid and Vapor Phase Tellurium,” ACS Catal, vol. 10, no. 15, pp. 8223–
8230, Aug. 2020, doi:
10.1021/ACSCATAL.0C00805/ASSET/IMAGES/LARGE/CS0C00805_0008.JPEG.

[71] R. Dagle, V. Dagle, M. Bearden, J. Holladay, T. Krause, and S. Ahmed, “An Overview of Natural Gas
Conversion Technologies for Co-Production of Hydrogen and Value-Added Solid Carbon
Products,” 2017. [Online]. Available: https://www.congress.gov/114/crpt/srpt236/CRPT-
114srpt236.pdf,

[72] T. Keipi, H. Tolvanen, and J. Konttinen, “Economic analysis of hydrogen production by methane
thermal decomposition: Comparison to competing technologies,” Energy Convers Manag, vol.
159, pp. 264–273, Mar. 2018, doi: 10.1016/J.ENCONMAN.2017.12.063.

[73] M. Msheik, S. Rodat, and S. Abanades, “Methane Cracking for Hydrogen Production: A Review of
Catalytic and Molten Media Pyrolysis,” Energies (Basel), vol. 14, no. 11, 2021, doi:
10.3390/en14113107.

[74] B. Gaudernack and S. Lynum, “Hydrogen from natural gas without release of CO2 to the
atmosphere,” Int J Hydrogen Energy, vol. 23, no. 12, pp. 1087–1093, Dec. 1998, doi:
10.1016/S0360-3199(98)00004-4.

[75] T. Keipi, V. Hankalin, J. Nummelin, and R. Raiko, “Techno-economic analysis of four concepts for
thermal decomposition of methane: Reduction of CO2 emissions in natural gas combustion,”
Energy Convers Manag, vol. 110, pp. 1–12, Feb. 2016, doi: 10.1016/J.ENCONMAN.2015.11.057.

[76] L. Fulcheri, N. Probst, G. Flamant, F. Fabry, E. Grivei, and X. Bourrat, “Plasma processing: a step
towards the production of new grades of carbon black,” Carbon N Y, vol. 40, no. 2, pp. 169–176,
Feb. 2002, doi: 10.1016/S0008-6223(01)00169-5.

[77] A. Mašláni et al., “Pyrolysis of methane via thermal steam plasma for the production of hydrogen
and carbon black,” Int J Hydrogen Energy, vol. 46, no. 2, pp. 1605–1614, Jan. 2021, doi:
10.1016/J.IJHYDENE.2020.10.105.

[78] L. Fulcheri, V.-J. Rohani, E. Wyse, N. Hardman, and E. Dames, “An energy-efficient plasma
methane pyrolysis process for high yields of carbon black and hydrogen,” Int J Hydrogen Energy,
Nov. 2022, doi: 10.1016/J.IJHYDENE.2022.10.144.

[79] J. J. Moss and B. T. Noel, “DC plasma torch electrical power design method and apparatus,”
20170034898, Feb. 02, 2017

[80] K. S. Kim, S. H. Hong, K. S. Lee, and W. T. Ju, “Continuous synthesis of nanostructured sheetlike
carbons by thermal plasma decomposition of methane,” IEEE Transactions on Plasma Science,
vol. 35, no. 2 III, pp. 434–443, Apr. 2007, doi: 10.1109/TPS.2007.892556.

78
[81] L. Zhang et al., “Visible-light-driven non-oxidative dehydrogenation of alkanes at ambient
conditions,” Nature Energy 2022, pp. 1–10, Sep. 2022, doi: 10.1038/s41560-022-01127-1.

[82] A. Abánades, C. Rubbia, and D. Salmieri, “Technological challenges for industrial development of
hydrogen production based on methane cracking,” Energy, vol. 46, no. 1, pp. 359–363, Oct. 2012,
doi: 10.1016/J.ENERGY.2012.08.015.

[83] M. Steinberg, “Fossil fuel decarbonization technology for mitigating global warming,” Int J
Hydrogen Energy, vol. 24, no. 8, pp. 771–777, Aug. 1999, doi: 10.1016/S0360-3199(98)00128-1.

[84] N. Z. Muradov, “How to produce hydrogen from fossil fuels without CO2 emission,” Int J
Hydrogen Energy, vol. 18, no. 3, pp. 211–215, Mar. 1993, doi: 10.1016/0360-3199(93)90021-2.

[85] N. Shah, D. Panjala, and G. P. Huffman, “Hydrogen production by catalytic decomposition of


methane,” Energy and Fuels, vol. 15, no. 6, pp. 1528–1534, Nov. 2001, doi: 10.1021/EF0101964.

[86] N. S. N. Hasnan, S. N. Timmiati, K. L. Lim, Z. Yaakob, N. H. N. Kamaruddin, and L. P. Teh, “Recent


developments in methane decomposition over heterogeneous catalysts: an overview,” Mater
Renew Sustain Energy, vol. 1, Apr. 2020, doi: 10.1007/s40243-020-00167-5.

[87] N. Muradov, “Low to near-zero CO2 production of hydrogen from fossil fuels: Status and
perspectives,” Int J Hydrogen Energy, vol. 42, no. 20, pp. 14058–14088, May 2017, doi:
10.1016/J.IJHYDENE.2017.04.101.

[88] N. Bayat, F. Meshkani, and M. Rezaei, “Thermocatalytic decomposition of methane to COx-free


hydrogen and carbon over Ni–Fe–Cu/Al2O3 catalysts,” Int J Hydrogen Energy, vol. 41, no. 30, pp.
13039–13049, Aug. 2016, doi: 10.1016/J.IJHYDENE.2016.05.230.

[89] C. H. Bartholomew, “Mechanisms of catalyst deactivation,” Appl Catal A Gen, vol. 212, no. 1–2,
pp. 17–60, Apr. 2001, doi: 10.1016/S0926-860X(00)00843-7.

[90] M. G. Poirier and C. Sapundzhiev, “Catalytic decomposition of natural gas to hydrogen for fuel
cell applications,” Int J Hydrogen Energy, vol. 22, no. 4, pp. 429–433, Apr. 1997, doi:
10.1016/S0360-3199(96)00101-2.

[91] R. Aiello, J. E. Fiscus, H. C. zur Loye, and M. D. Amiridis, “Hydrogen production via the direct
cracking of methane over Ni/SiO2: catalyst deactivation and regeneration,” Appl Catal A Gen, vol.
192, no. 2, pp. 227–234, Feb. 2000, doi: 10.1016/S0926-860X(99)00345-2.

[92] N. Muradov, “Catalysis of methane decomposition over elemental carbon,” Catal Commun, vol.
2, no. 3–4, pp. 89–94, Jul. 2001, doi: 10.1016/S1566-7367(01)00013-9.

[93] N. Muradov, F. Smith, and A. T-Raissi, “Catalytic activity of carbons for methane decomposition
reaction,” Catal Today, vol. 102–103, pp. 225–233, May 2005, doi:
10.1016/J.CATTOD.2005.02.018.

[94] N. Z. Muradov, “CO2-free production of hydrogen by catalytic pyrolysis of hydrocarbon fuel,”


Energy and Fuels, vol. 12, no. 1, pp. 41–48, 1998, doi:
10.1021/EF9701145/ASSET/IMAGES/LARGE/EF9701145F00013.JPEG.

79
[95] D. P. Serrano, J. A. Botas, and R. Guil-Lopez, “H2 production from methane pyrolysis over
commercial carbon catalysts: Kinetic and deactivation study,” Int J Hydrogen Energy, vol. 34, no.
10, pp. 4488–4494, May 2009, doi: 10.1016/J.IJHYDENE.2008.07.079.

[96] U. P. M. Ashik, W. M. A. Wan Daud, and H. F. Abbas, “Production of greenhouse gas free
hydrogen by thermocatalytic decomposition of methane – A review,” Renewable and Sustainable
Energy Reviews, vol. 44, pp. 221–256, Apr. 2015, doi: 10.1016/J.RSER.2014.12.025.

[97] Z. Fan, W. Weng, J. Zhou, D. Gu, and W. Xiao, “Catalytic decomposition of methane to produce
hydrogen: A review,” Journal of Energy Chemistry, vol. 58, pp. 415–430, Jul. 2021, doi:
10.1016/J.JECHEM.2020.10.049.

[98] N. Muradov, Z. Chen, and F. Smith, “Fossil hydrogen with reduced CO2 emission: Modeling
thermocatalytic decomposition of methane in a fluidized bed of carbon particles,” Int J Hydrogen
Energy, vol. 30, no. 10, pp. 1149–1158, Aug. 2005, doi: 10.1016/J.IJHYDENE.2005.04.005.

[99] D. P. Serrano, J. Á. Botas, P. Pizarro, and G. Gómez, “Kinetic and autocatalytic effects during the
hydrogen production by methane decomposition over carbonaceous catalysts,” Int J Hydrogen
Energy, vol. 38, no. 14, pp. 5671–5683, May 2013, doi: 10.1016/J.IJHYDENE.2013.02.112.

[100] A. Domínguez, B. Fidalgo, Y. Fernández, J. J. Pis, and J. A. Menéndez, “Microwave-assisted


catalytic decomposition of methane over activated carbon for CO2-free hydrogen production,”
Int J Hydrogen Energy, vol. 32, no. 18, pp. 4792–4799, Dec. 2007, doi:
10.1016/J.IJHYDENE.2007.07.041.

[101] National Aeronautics and Space Administration and Science Mission Directorate, “Introduction to
the Electromagnetic Spectrum,” 2010. https://science.nasa.gov/ems/01_intro (accessed Nov. 15,
2022).

[102] L. Zong, S. Zhou, N. Sgriccia, M. C. Hawley, and L. C. Kempel, “A Review of Microwave-Assist


Polymer Chemistry (MAPC),” http://dx.doi.org/10.1080/08327823.2003.11688487, vol. 38, no. 1,
pp. 49–74, 2016, doi: 10.1080/08327823.2003.11688487.

[103] T. Kim, J. Lee, and K. H. Lee, “Microwave heating of carbon-based solid materials,” Carbon
Letters, vol. 15, no. 1, pp. 15–24, Jan. 2014, doi: 10.5714/CL.2014.15.1.015.

[104] E. T. Thostenson and T. W. Chou, “Microwave processing: fundamentals and applications,”


Compos Part A Appl Sci Manuf, vol. 30, no. 9, pp. 1055–1071, Sep. 1999, doi: 10.1016/S1359-
835X(99)00020-2.

[105] J. R. Grace, J. Chaouki, and T. Pugsley, “Fluidized Bed Reactor,” in Particle Technology and
Applications, CRC Press, 2017, pp. 369–403. doi: 10.1016/b978-0-12-410416-7.00008-2.

[106] R. Cocco, S. B. Reddy, and K. T. Knowlton, “Back to Basics Introduction to Fluidization,” 2014,
Accessed: Nov. 17, 2022. [Online]. Available: www.aiche.org/cep

[107] J. B. Cheng, H. G. Shi, M. Cao, T. Wang, H. B. Zhao, and Y. Z. Wang, “Porous carbon materials for
microwave absorption,” Mater Adv, vol. 1, no. 8, pp. 2631–2645, Nov. 2020, doi:
10.1039/D0MA00662A.

80
[108] H. F. Abbas and W. M. A. Wan Daud, “Hydrogen production by methane decomposition: A
review,” Int J Hydrogen Energy, vol. 35, no. 3, pp. 1160–1190, Feb. 2010, doi:
10.1016/J.IJHYDENE.2009.11.036.

[109] K. Zuraiqi et al., “Liquid Metals in Catalysis for Energy Applications,” Joule, vol. 4, no. 11, pp.
2290–2321, Nov. 2020, doi: 10.1016/J.JOULE.2020.10.012.

[110] A. Abánades et al., “Development of methane decarbonisation based on liquid metal technology
for CO2-free production of hydrogen,” Int J Hydrogen Energy, vol. 41, no. 19, pp. 8159–8167,
May 2016, doi: 10.1016/J.IJHYDENE.2015.11.164.

[111] S. Timmerberg, M. Kaltschmitt, and M. Finkbeiner, “Hydrogen and hydrogen-derived fuels


through methane decomposition of natural gas – GHG emissions and costs,” Energy Conversion
and Management: X, vol. 7, p. 100043, Sep. 2020, doi: 10.1016/J.ECMX.2020.100043.

[112] N. A. Juan, A. Naseri, M. R. Kholghy, and M. J. Thomson, “NanoParticle Flow Reactor (NanoPFR): a
tested model for simulating carbon nanoparticle formation in flow reactors,” International
Journal of Chemical Reactor Engineering, Aug. 2022, doi: 10.1515/IJCRE-2021-
0258/ASSET/GRAPHIC/J_IJCRE-2021-0258_FIG_008.JPG.

[113] G. Blanquart, P. Pepiot-Desjardins, and H. Pitsch, “Chemical mechanism for high temperature
combustion of engine relevant fuels with emphasis on soot precursors,” Combust Flame, vol.
156, no. 3, pp. 588–607, Mar. 2009, doi: 10.1016/J.COMBUSTFLAME.2008.12.007.

[114] S. H. Park and S. N. Rogak, “A novel fixed-sectional model for the formation and growth of
aerosol agglomerates,” J Aerosol Sci, vol. 35, no. 11, pp. 1385–1404, Nov. 2004, doi:
10.1016/J.JAEROSCI.2004.05.010.

[115] Jin Jwang Wu and R. C. Flagan, “A discrete-sectional solution to the aerosol dynamic equation,” J
Colloid Interface Sci, vol. 123, no. 2, pp. 339–352, Jun. 1988, doi: 10.1016/0021-9797(88)90255-X.

[116] M. v. Smoluchowski, “Versuch einer mathematischen Theorie der Koagulationskinetik kolloider


Lösungen,” Zeitschrift für Physikalische Chemie, vol. 92U, no. 1, pp. 129–168, Nov. 1918, doi:
10.1515/ZPCH-1918-9209.

[117] D. D. Tanner, P. Kandanarachchi, Q. Ding, H. Shao, D. Vizitiu, and J. A. Franz, “The catalytic
conversion of C1-Cn hydrocarbons to olefins and hydrogen: Microwave-assisted C-C and C-H
bond activation,” Energy and Fuels, vol. 15, no. 1, pp. 197–204, 2001, doi:
10.1021/EF000167D/ASSET/IMAGES/LARGE/EF000167DF00002.JPEG.

[118] M. Frenklach, D. W. Clary, W. C. Gardiner, and S. E. Stein, “Detailed kinetic modeling of soot
formation in shock-tube pyrolysis of acetylene,” Symposium (International) on Combustion, vol.
20, no. 1, pp. 887–901, Jan. 1985, doi: 10.1016/S0082-0784(85)80578-6.

[119] M. Frenklach and H. Wang, “Detailed modeling of soot particle nucleation and growth,”
Symposium (International) on Combustion, vol. 23, no. 1, pp. 1559–1566, Jan. 1991, doi:
10.1016/S0082-0784(06)80426-1.

81
[120] T. Becker, M. Richter, and D. W. Agar, “Methane pyrolysis: Kinetic studies and mechanical
removal of carbon deposits in reactors of different materials,” Int J Hydrogen Energy, Nov. 2022,
doi: 10.1016/J.IJHYDENE.2022.10.069.

[121] J. F. Parker et al., “Pyrolytic Carbon Films with Tunable Electronic Structure and Surface
Functionality: A Planar Stand-In for Electroanalysis of Energy-Relevant Reactions,”
ChemElectroChem, vol. 7, no. 3, pp. 672–683, Feb. 2020, doi: 10.1002/CELC.201901672.

[122] W. Deng, Y. Su, S. Liu, and H. Shen, “Microwave-assisted methane decomposition over pyrolysis
residue of sewage sludge for hydrogen production,” Int J Hydrogen Energy, vol. 39, no. 17, pp.
9169–9179, Jun. 2014, doi: 10.1016/J.IJHYDENE.2014.04.033.

[123] A. Holmen, O. Olsvik, and O. A. Rokstad, “Pyrolysis of natural gas: chemistry and process
concepts,” Fuel Processing Technology, vol. 42, no. 2–3, pp. 249–267, Apr. 1995, doi:
10.1016/0378-3820(94)00109-7.

[124] S. Patel et al., “Production of hydrogen by catalytic methane decomposition using biochar and
activated char produced from biosolids pyrolysis,” Int J Hydrogen Energy, vol. 45, no. 55, pp.
29978–29992, Nov. 2020, doi: 10.1016/J.IJHYDENE.2020.08.036.

[125] Y. Luo, Y. Shi, and N. Cai, “Bridging a bi-directional connection between electricity and fuels in
hybrid multienergy systems,” in Hybrid Systems and Multi-energy Networks for the Future Energy
Internet, Elsevier, 2021, pp. 41–84. doi: 10.1016/b978-0-12-819184-2.00003-1.

[126] C. Coutanceau, S. Baranton, and T. Audichon, “Hydrogen Production From Water Electrolysis,”
Hydrogen Electrochemical Production, pp. 17–62, 2018, doi: 10.1016/B978-0-12-811250-2.00003-
0.

[127] A. S. Ansar, A. S. Gago, F. Razmjooei, R. Reißner, Z. Xu, and K. A. Friedrich, “Alkaline electrolysis—
status and prospects,” Electrochemical Power Sources: Fundamentals, Systems, and Applications,
pp. 165–198, 2022, doi: 10.1016/B978-0-12-819424-9.00004-5.

[128] J. Ivy, “Summary of Electrolytic Hydrogen Production: Milestone Completion Report,” 2003,
Accessed: Oct. 09, 2022. [Online]. Available: http://www.osti.gov/bridge

[129] M. Carmo, D. L. Fritz, J. Mergel, and D. Stolten, “A comprehensive review on PEM water
electrolysis,” Int J Hydrogen Energy, vol. 38, no. 12, pp. 4901–4934, Apr. 2013, doi:
10.1016/J.IJHYDENE.2013.01.151.

[130] P. Millet, D. Dragoe, S. Grigoriev, V. Fateev, and C. Etievant, “GenHyPEM: A research program on
PEM water electrolysis supported by the European Commission,” Int J Hydrogen Energy, vol. 34,
no. 11, pp. 4974–4982, Jun. 2009, doi: 10.1016/J.IJHYDENE.2008.11.114.

[131] P. Millet, “PEM Water Electrolysis,” Hydrogen Production: By Electrolysis, pp. 63–116, Feb. 2015,
doi: 10.1002/9783527676507.ch3.

[132] S. A. Grigoriev, V. I. Porembsky, and V. N. Fateev, “Pure hydrogen production by PEM electrolysis
for hydrogen energy,” Int J Hydrogen Energy, vol. 31, no. 2, pp. 171–175, Feb. 2006, doi:
10.1016/J.IJHYDENE.2005.04.038.

82
[133] S. Shiva Kumar and V. Himabindu, “Hydrogen production by PEM water electrolysis – A review,”
Mater Sci Energy Technol, vol. 2, no. 3, pp. 442–454, Dec. 2019, doi:
10.1016/J.MSET.2019.03.002.

[134] M. S. Thomassen, A. H. Reksten, A. O. Barnett, T. Khoza, and K. Ayers, “PEM water electrolysis,”
Electrochemical Power Sources: Fundamentals, Systems, and Applications, pp. 199–228, 2022,
doi: 10.1016/B978-0-12-819424-9.00013-6.

[135] F. Barbir, “PEM electrolysis for production of hydrogen from renewable energy sources,” Solar
Energy, vol. 78, no. 5, pp. 661–669, May 2005, doi: 10.1016/J.SOLENER.2004.09.003.

[136] M. Kopp, D. Coleman, C. Stiller, K. Scheffer, J. Aichinger, and B. Scheppat, “Energiepark Mainz:
Technical and economic analysis of the worldwide largest Power-to-Gas plant with PEM
electrolysis,” Int J Hydrogen Energy, vol. 42, no. 19, pp. 13311–13320, May 2017, doi:
10.1016/J.IJHYDENE.2016.12.145.

[137] J. Chi and H. Yu, “Water electrolysis based on renewable energy for hydrogen production,”
Chinese Journal of Catalysis, vol. 39, no. 3, pp. 390–394, Mar. 2018, doi: 10.1016/S1872-
2067(17)62949-8.

[138] S. A. Grigoriev and V. N. Fateev, “Hydrogen Production by Water Electrolysis,” Hydrogen


Production Technologies, pp. 231–276, Mar. 2017, doi: 10.1002/9781119283676.CH6.

[139] M. Zahid, J. Schefold, and A. Brisse, “High-Temperature Water Electrolysis Using Planar Solid
Oxide Fuel Cell Technology: a Review,” Essen Schriften des Forschungszentrums Jülich / Energy &
Environment, vol. 78, 2010.

[140] M. Seitz, H. von Storch, A. Nechache, and D. Bauer, “Techno economic design of a solid oxide
electrolysis system with solar thermal steam supply and thermal energy storage for the
generation of renewable hydrogen,” Int J Hydrogen Energy, vol. 42, no. 42, pp. 26192–26202,
Oct. 2017, doi: 10.1016/J.IJHYDENE.2017.08.192.

[141] F. R. Bianchi, B. Bosio, F. José, and H. Fernández, “Operating Principles, Performance and
Technology Readiness Level of Reversible Solid Oxide Cells,” Sustainability 2021, Vol. 13, Page
4777, vol. 13, no. 9, p. 4777, Apr. 2021, doi: 10.3390/SU13094777.

[142] M. Liang, B. Yu, M. Wen, J. Chen, J. Xu, and Y. Zhai, “Preparation of LSM–YSZ composite powder
for anode of solid oxide electrolysis cell and its activation mechanism,” J Power Sources, vol. 190,
no. 2, pp. 341–345, May 2009, doi: 10.1016/J.JPOWSOUR.2008.12.132.

[143] O. Schmidt, A. Gambhir, I. Staffell, A. Hawkes, J. Nelson, and S. Few, “Future cost and
performance of water electrolysis: An expert elicitation study,” Int J Hydrogen Energy, vol. 42, no.
52, pp. 30470–30492, Dec. 2017, doi: 10.1016/J.IJHYDENE.2017.10.045.

[144] A. Nechache and S. Hody, “Alternative and innovative solid oxide electrolysis cell materials: A
short review,” Renewable and Sustainable Energy Reviews, vol. 149, p. 111322, Oct. 2021, doi:
10.1016/J.RSER.2021.111322.

[145] K. Chen and S. P. Jiang, “Review—Materials Degradation of Solid Oxide Electrolysis Cells,” J
Electrochem Soc, vol. 163, no. 11, pp. F3070–F3083, Jun. 2016, doi: 10.1149/2.0101611JES/XML.

83
[146] G. Schiller, A. Ansar, M. Lang, and O. Patz, “High temperature water electrolysis using metal
supported solid oxide electrolyser cells (SOEC),” J Appl Electrochem, vol. 39, no. 2, pp. 293–301,
Feb. 2009, doi: 10.1007/S10800-008-9672-6/FIGURES/12.

[147] J. Appel, H. Bockhorn, and M. Frenklach, “Kinetic modeling of soot formation with detailed
chemistry and physics: laminar premixed flames of C2 hydrocarbons,” Combust Flame, vol. 121,
no. 1–2, pp. 122–136, Apr. 2000, doi: 10.1016/S0010-2180(99)00135-2.

[148] K. Gleason, F. Carbone, A. J. Sumner, B. D. Drollette, D. L. Plata, and A. Gomez, “Small aromatic
hydrocarbons control the onset of soot nucleation,” Combust Flame, vol. 223, pp. 398–406, Jan.
2021, doi: 10.1016/J.COMBUSTFLAME.2020.08.029.

84
Appendix A Electrolysis Technologies
A.1 Alkaline Water Electrolysis (AWE)
Alkaline water electrolysis (AWE) is the oldest and most mature electrolysis technology used
today capable of reaching megawatt (MW) capacity plants lasting about 7 – 15 years [47], [48], [125].
AWE consists of two electrodes, typically nickel, submersed in an aqueous solution of about 25 – 30
wt.% potassium hydroxide (KOH) or sodium hydroxide (NaOH) as a liquid electrolyte to allow ionic
conductivity [48], [125], [126]. KOH has become the standard solution for AWE [127]. A gas-tight
diaphragm is placed in between the two electrodes to prevent the gases from crossing and recombining.
The diaphragm is constructed of an inorganic ion-exchange-type membrane that is pressed up closely by
the electrodes to minimize the space between them and reduce ohmic resistance [46], [48]. The typical
operating temperature for this setup is anywhere between 5° - 100°C [46], [48]. This is illustrated in
Figure A1 below showing the hydroxyl ions passing through the diaphragm towards the anode and make
oxygen and hydrogen [46]. The reaction equations for this type of electrolysis process (AWE) are the
same as in equations (17) & (18).

Figure A1. Operating principle of alkaline water electrolysis [46]

This method can produce hydrogen at a purity of at least 99.7% [48], [128]. They are suitable for
large-scale hydrogen production able to output 500 – 760 Nm3/h which corresponds to about 2150 – 3534
kW of power consumption [46]. Since this technology is more established and uses cheaper metals, such
as nickel, it has been able to achieve lower costs compared to PEME and SOE [129]. The drawbacks for
this method is that the corrosion of electrodes from the alkaline electrolyte and a lower efficiency (60 –
82%) compared to PEM and SOE [47], [129].

Research activity in this process is focused on optimization of electrodes/catalysts, diaphragm


material, current density and minimizing space in between the electrodes to reduce operating costs
associated with electricity consumption and improved durability and efficiency [46], [48], [127], [129].

85
A.2 Proton Exchange Membrane Electrolysis (PEM)
The first proton exchange membrane (PEM) was developed in 1966 by General Electric Co. to
combat the negative aspects that came with alkaline water electrolysis [48], [130]. The electrolyte in this
technology involves a thin gas-tight polymeric membrane of about 0.2 mm capable of conducting protons
(H+) through an ion exchange mechanism [46], [131]. These cells become compact allowing for water
splitting efficiency to be high [131]. In addition to conducting protons through it, the membrane prevents
recombination of molecular hydrogen and oxygen back into water [131]. The most common material used
as a membrane is Nafion [46], [131]. The membrane is placed in between two electrodes typically made
of noble metals such as platinum (Pt) and iridium (Ir) [132]. The membranes used are not 100% gas proof
so there will be some crossover in the gaseous species [48]. PEM electrolysis typically operates at 20 –
80°C and 30 bar [133], [134]. If the pressure were to increase, the rate of permeation of gases through the
membrane may increase due to Fick’s law of diffusion.

Water is typically fed through the anode where DC current will oxidize the water into oxygen
(O2), protons (H+), and electrons (e-) as seen in equation (28). In response to the induced electric field,
the protons will travel through the ion conducting membrane towards the cathode according to Figure A2
where they reduce into molecular hydrogen as shown in equation (29) [46].

Figure A2. Operating principle of proton exchange membrane electrolysis cell [46]

1
Anode: 𝐻2 𝑂(𝑙) → 𝑂2 (𝑔) + 2𝐻 + (𝑎𝑞) + 2𝑒 − (28)
2

Cathode: 2𝐻 + (𝑎𝑞) + 2𝑒 − → 𝐻2 (𝑔) (29)

This provides a hydrogen purity of >99.9% and even up to >99.999% in some cases [135].
Implementation of PEM electrolysis at the megawatt scale is possible but it is more expensive compared
to AWE and are thus only available for low-scale production [48]. Low-scale production applications
typically only produce hydrogen at a rate of 30 Nm3/h with an associated power consumption of 174 kW
[46]. The high investment costs are usually due to the costly noble metal-based electrodes, PEM has a
shorter lifetime than alkaline technology and lower hydrogen production capacity [46].

86
Although AWE is currently cheaper and reliable than PEM electrolysis, the latter has a stronger
response time in starting up compared to AWE. In addition to investment costs becoming cheaper
overtime relative to AWE, PEM could be a better choice for intermittent hydrogen production from
excess renewable energy [136], [137].

A.3 Solid Oxide Electrolysis (SOEC)


From a thermodynamics perspective, splitting water at elevated temperatures is more attractive as
the energy required to split water decreases when the cell temperature increases [138]. Solid oxide
electrolysis (SOEC) operates at around 500 – 1000°C with efficiencies ranging from at least 80% to
almost 100% [47], [129], [139]–[141]. SOEC is in the R&D stage and is the least developed technology
compared to AWE and PEM and are not yet commercialized [48].

The basic operating principle is similar to that of any other electrolysis technology but with a
different electrolyte and high temperature steam in place of water in SOEC. As shown in Figure A3, high
temperature steam is supplied to the cathode side where it is reduced to hydrogen and oxygen ions as seen
in equation (30) [46], [47]. The oxygen ions pass through solid electrolyte membrane where they
recombine to form oxygen gas in equation (31). The membranes are oxygen ion conductors typically
made from nickel/yttria stabilized zirconia [48], [142].

Figure A3. Operating principle of solid oxide electrolysis cell [46]

Cathode: 𝐻2 𝑂(𝑔) + 2𝑒 − → 𝐻2 (𝑔) + 𝑂2− (30)

1
Anode: 𝑂2− → 2 𝑂2 (𝑔) + 2𝑒 − (31)

As previously mentioned, this process is highly efficient but still only in the R&D phase as it has
its host of challenges. The common problems of SOEC technology are rapid cell degradation, electrode
corrosion, long turn-on and turn-off time and shorter lifespan [48], [143]–[146]. It is believed that there is
potential to produce large amounts of hydrogen with this technology if the aforementioned problems
related to durability at high operating temperatures can be solved [129].

87
Appendix B Experiment Setup Procedure
1. Drop at least three fritted discs into a clean quartz tube. Ensure that the discs are level as per
Figure B1 and not tilted.

Figure B1

2. Pour desired amount of carbon (40 – 80g) into quartz tube.


3. Head over to the microwave with the carbon filled quartz tube. Line up the muffle with the
chokes, slide quartz tube into the choke and muffle.
4. On the bottom end, slide on ultra torr fitting as per Figure B2.

Figure B2

5. Attach inlet flow line connector to the bottom quartz end and fasten ultra torr fitting onto inlet
flow connection as per Figure B3.

88
Figure B3

6. For the top end of the quartz, repeat Step 4 but in reverse except without the ultra torr nut as per
Figure B4.

Figure B4

7. Fasten the tee piece with thermocouple attached to the top choke.
8. Connect thermocouple’s male connector to its female counterpart.
9. Place muffle door on muffle.
10. Disassemble paper filter housing and replace old filter paper with a new filter paper. Reassemble
in reverse order.

89
Appendix C Experiment Operation Procedure
This procedure assumes the experiment (carbon in quartz, new filter paper, muffle etc.) has been set up.
1. Turn ON microwave and flow controllers
2. Open nitrogen cylinder valve. Ensure the pressure in the pressure regulators matches the inlet
pressure of the flow controller
3. Set nitrogen flow rate on the controller to the desired flow rate (typically 0.2 L/min). See flow
calibration procedure in Appendix F to obtain accurate flow rates.
4. Flush carbon for about 5 minutes with nitrogen.
5. Ensure ventilation fan plug is connected
6. Open LabVIEW script for measuring temperature.
7. Double click on ‘Write to Measurement File’ and change the date to date of experiment. This will
create a folder titled with the date and write the temperature and time in a text file.
8. Click ‘Run’ to start measuring temperature.
9. Turn ‘Microwave Power’ ON.
10. Adjust dial to desired power level. Note that the numbers on the power dial do not correlate to the
percentage of output power of the microwave’s capacity.
11. Set timer to desired microwave run time.
12. Press the “Timer Reset” button to start the timer.
13. Observe the temperature and ensure it is rising.
14. Turn ON GC-TCD following the GC-TCD operation instructions in Appendix D.
15. Once temperature rises to the desired level, open methane cylinder.
16. Set methane flow rate while also adjusting nitrogen flow rate to obtain appropriate gas
composition. For example, if 50-50 methane to nitrogen is desired, lower nitrogen from 0.2 L/min
to 0.1 L/min and increase methane from 0 L/min to 0.1 L/min.
17. Run GC-TCD following instructions outlined in Appendix D
Shutdown Procedure:
1. Adjust power level dial on microwave to zero.
2. Turn ‘Microwave Power’ OFF.
3. Lower methane flow rate to 0 L/min while continuing to flow nitrogen. It is important for
nitrogen to continue flowing to prevent air from entering the microwave.
4. Prop open microwave door to allow increased circulation
5. Remove muffle door.
6. When temperature drops below 80°C, remove reactor.
7. Stop recording temperature.
8. Shutdown GC-TCD as per Appendix D.
9. Close all cylinder valves if no more experiments will be completed in the day.

90
Appendix D GC-TCD Operation Procedure
Operating Procedure:
1. Turn GC ON using button on the bottom
2. On the small screen, select ‘Login’
3. Wait 5 – 10 minutes
4. Turn air pressure on at 60 psi
5. Open TC Nav software on desktop home screen
6. Select ‘manager’ as username; password: 12345678 (subject to change)
7. Confirm the Ethernet cable is connected from the computer to the side of the GC.
8. Left click on ‘Run’ and select ‘Take Control’. This action allows the GC to communicate with the
software and make runs from the computer.
9. Select ‘Method’ under ‘Build’
10. Select ‘C:\GC Arnel\methods\NARL8552 MODEL 4016 AR PROCESSING METHOD’

Figure D1

11. Click ‘OK’


12. Select ‘Setup’ at the top.

Figure D2

91
13. The ‘Setup Instrument’ window will appear. In the ‘Base file name’ section, write the name of the
file in the format ‘YYMMDD-A-GCB’ where Y is year, M is month, D is day, A is the
experiment # and B is the run number. A will only change if an experimental variable has
changed such as temperature, mass, concentration, particle size, flow rate etc. B will change for
each run you do within the set experimental parameters. For example, 220627-1-GC3 would be
experiment 1, run 3 on June 27th, 2022.
14. Select ‘Suppress Reports/Plots’. This step is crucial in order for the software to not return an error
when trying to view the data.
15. Select ‘Vial list’

Figure D3

16. The ‘Sequence Editor – Vial List’ screen will come up (see Figure D4). This is where you can
load up multiple runs. The number of rows indicates the number of samples the GC will measure.
Note that after the allotted number of samples (number of rows) have been measured, the GC will
start to cool down to standard settings. Therefore, it is always good practice to have more rows
than needed. Initially, when first loading the ‘Sequence Editor – Vial List’, there will be only five
partially filled rows with no information in the ‘Name’ and ‘Number’ columns.

92
Figure D4

17. To add more rows, press ‘Ctrl + A’ or go to ‘Edit’ and select ‘Append’.
18. Fill out the ‘Name’ similar to the ‘Base file name’. Then press ‘Ctrl + D’ or select ‘Change’ and
‘Fill Down’ to fill the rest of the column.
19. Select the first empty cell in the ‘Number’ column, click ‘Change’ then select ‘Smart Fill…’.
20. ‘Starting row’ is 1, ‘Ending row’ is the number of rows that have been made in Step 17.
21. ‘Sample number pattern’ is ‘##’.
22. ‘Starting number’ is 1.
23. Click ‘OK’
24. ‘Save’ and exit the editor
25. Click ‘OK’ on ‘Setup Instrument’ window.
26. Wait for the temperature to reach the set temperature (200°C) and for the GC screen to go from
‘Not Ready’ (red text) to ‘Ready’ (green text). There are three lines of text that are required to
display ‘Ready’ in order to proceed with the experiment. See Figure D5 below.

93
Figure D5

27. Once everything in Figure D5 display ready in green text, click ‘Run’ and select ‘Start Run’.
When you do that, you will hear gas sounds coming from the GC.
28. You can click ‘Real-Time Plot’ under ‘View’ to see the sample be detected live. The entire
process will take roughly 11 minutes.
29. Once it is done collecting, select ‘Results’ under ‘Reprocess’ and find the file you recently
named. Select ‘Display’ and then ‘Peak Report’. The following page should appear as Figure D6
below. The values of each detected species can be selected by clicking on their peaks and their
corresponding values will be highlighted in blue. In Figure D6, the methane peak is selected and
the corresponding values are highlighted in blue.

94
Figure D6

95
Shutdown Procedure:
1. Select ‘Method’ under ‘Build’
2. Select ‘Shutdown’
3. Click ‘OK’
4. Select ‘Setup’
5. Name the file with the date and labelled shutdown e.g. YYMMDD-shutdown
6. Wait until temperature reaches 100°C and the status is the same as in Figure D5 (this may take a
while)
7. Click ‘Run’
8. Select ‘Release Control’
9. Click ‘Run’ and then select ‘Clear Setup’
10. Once the status becomes red (red text), press the power button on the side of the GC to turn it off
11. Close the air valve

96
Appendix E Fused Quartz Tube Properties
The properties of the quartz tube below are obtained on the website of the manufacturer,
Technical Glass Products, Inc in Fall of 2022.
Table 5. Quartz tube properties from Technical Glass Products, Inc.
PROPERTY TYPICAL VALUES
Density 2.2 x 103 kg/m3
Hardness 5.5–6.5 Mohs’s Scale
570 KHN100
Design Tensile Strength 4.8 x 107 Pa (N/m2 ) (7,000 psi)
Design Compressive Strength >1.1 x 109 Pa (160,000 psi)
Bulk Modulus 3.7 x 101 0 Pa (5.3 x 106 psi)
Rigidity Modulus 3.1 x 101 0 Pa (4.5 x 106 psi)
Young’s Modulus 7.2 x 101 0 Pa (10.5 x 106 psi)
Poisson’s Ratio 0.17
Coefficient of Thermal Expansion (20°C– 5.5 x 10-7 cm/cm °C
320°C)
Thermal Conductivity (20°C) 1.4 W/m °C
Specific Heat (20°) 670 J/kg °C
Softening Point 1683°C
Annealing Point 1215°C
Strain Point 1120°C
Electrical Resistivity 7 x 107 ohm cm at 350°C
Dielectric Loss Factor <0.0004 at 20°C and 1 MHz
Dielectric Constant 3.75 at 20°C and 1 MHz
Dielectric Strength 5 x 107 V/m at 20°C and 1 MHz
Dissipation Factor <0.0001 at 20°C and 1 MHz
Index of Refraction 1.4585
Constringence (Nu value) 67.56
Velocity of Sound-Shear Wave 3.75 x 103 m/s
Velocity of Sound/Compressional Wave 5.90 x 103 m/s
Sonic Attenuation <11dB/m MHz
Permeability Constants (700°C)
Helium 2.1 x 10-8 cm3 mm/cm2 sec. cm Hg
Hydrogen 2.1 x 10-9 cm3 mm/cm2 sec. cm Hg
Deuterium 1.7 x 10-9 cm3 mm/cm2 sec. cm Hg
Neon 9.5 x 10-1 0 cm3 mm/cm2 sec. cm Hg

97
Appendix F Flow Controller Calibration Procedure
The mass flow controller and the control box must be calibrated for the respective gas that it will
control. The calibration procedure for the mass flow controllers is outlined below.
1. The mass flow controller (either Brooks or Bronkhorst) is connected to its respective control box.
The values on the control box display numbers with no units or meaning and do not show the real
flow rate hence the need for calibration.
2. To obtain the real flowrate, a ¼-inch tube is attached from the outlet of the mass flow controller
to a Bios DryCal Defender 530+ flow meter from MesaLabs. This device measures the real flow
rate that passes through it in mL/min.
3. To calibrate the mass flow controller, a calibration curve needs to be obtained.
4. The points on the curve is obtained by inputting values on the control box which will output the
corresponding real flow rate in mL/min on the Bios flow meter. The values in the Bios should be
recorded
5. The real flow rates are recorded for a number of values input in the control box so that a linear
calibration curve can be made. An example of this is tabulated below where the left column is the
values in the control box and the right column is its associated flow rate value. The numbers in
this table outputs equation (25).
Table 6. Table displaying values inputted in Brooks mass flow controller with resulting flow rate shown in Bios. From this, the
calibration curve in equation (25) was determined.
Mass Flow Control Box - Brooks Bios (L/min)
50 0.03515
100 0.07645
150 0.1178
200 0.15931
250 0.20047
300 0.24135

98
Appendix G SEM Images of Carbon Granules
Additional images on activated carbon particles before and after methane pyrolysis and thin film.

Figure G1. Activated carbon before methane pyrolysis at 50 micrometer scale

Figure G2. Activated carbon before methane pyrolysis at 30 micrometer scale

99
Figure G3. Activated carbon after methane pyrolysis at 50 micrometer scale

Figure G4. Activated carbon after methane pyrolysis at 30 micrometer scale

100
Figure G5. Thin film carbon at 300 micrometer scale

Figure G6. Thin film carbon at 50 micrometer scale

101
Appendix H Inception Model, Simulation Conditions and
Surface Growth
The inception model is developed based on the dimerization of PAHs, i.e., collision rate of two
PAHs to form a dimer. Based on the literature [147], [148], the model uses pyrene (C16) as the precursor
for carbon particle formation. The surface growth is modeled using two mechanisms: PAHs adsorption
and HACA. Similar to the inception model, the collision rate of PAHs with the carbon surface is the root
of PAH adsorption model. Carbon particles can also grow via heterogeneous reactions with light gas
phase hydrocarbons like acetylene on an active site. 60 sections are used to attain high resolution while
keeping the computational time reasonable. To calculate the mass of the first section, 400 carbon atoms
are selected as the incipient particles. In the sectional model, sections are divided into two parts, namely,
nanoparticles and pellets. The former corresponds to the formation and growth of carbon nanoparticles,
and the latter provides the information of carbon addition onto the surface of the initial particles. To
mimic the conditions of our experiment, the pellet section is initially filled with carbon particles with an
average diameter of 300 microns. A summary of the simulation conditions is provided below.
Table 7. Simulation conditions. The geometry, boundary conditions and sectional model parameters used in the NanoPFR code
are shown here

Geometry
Reactor length (cm) 25.4
Reactor diameter (cm) 2.25
Axial interval (cm) 10-4

Boundary conditions
Isothermal:
Temperature range (°C) [800, 1250]
Isobar:
Pressure (kPa) 101.325
Inlet fuel composition: (volume fraction)
Methane 0.5
Nitrogen 0.5
Inlet volumetric flow rate (L/min) 0.2
Gas-phase chemistry Caltech [113]

Sectional conditions
Number of sections 60
Spacing factor 1.9
Number of carbon atoms in the first section 400
3
Density of carbon particles (kg/m ) 1900
Pellet diameter (micron) 300
Pellet section number* 57
*This is the section where initial carbon particles are injected

102
Appendix I Effective Activation Energy Calculation
The mole balance equation for methane reacting along the reaction vessel is calculated below:
−dMCH4 + rCH4 dV = 0

where MCH4 , rCH4 , and V represent methane molar flow rate, methane consumption rate, and reactor
volume. The rate of methane consumption is −k [CH4 ] where [CH4 ] represents the concentration of
MCH4
methane and is equal to −k × Q
where Q is the methane volumetric flow rate, and k is the reaction rate
constant.
By substituting the methane consumption rate in the mole balance equation and rearranging, it gives:
Q MCH4 i
k= ln ( )
V MCH4 o

Above relation can be also written as a function of residence time (t r ) and methane conversion (X)
calculated blow:
V MCH4 o
tr = , X=1−
Q MCH4 i

Therefore, the final form of reaction rate constant is calculated as:


1
ln (1 − X)
k=
tr
It should be noted that t r is the residence time in the hot section of the reactor and it changes under
different reactor temperatures and flow rates. Afterwards, if we rearrange the following Arrhenius
equation, the value of activation energy (Ea ) can be obtained by calculating the slope of the line displayed
in Fig. 4.
−Ea
k = Aexp( )
RT
where A is the pre-exponential factor, R is the gas constant, and T is the temperature.

103

You might also like