Download as pdf or txt
Download as pdf or txt
You are on page 1of 54

Feasibility Analysis of Blending Hydrogen into Natural Gas Networks

Josmar B. Cristello1, Jaehyun M. Yang1, Simon S. Park1, Ron Hugo1, Youngsoo Lee2

1
University of Calgary, Canada
2
Jeonbuk National University, Jeonju, South Korea

Abstract
Hydrogen fuel has the potential to mitigate the negative effects of greenhouse gases and
climate change by neutralizing carbon emissions. Transporting large volume of hydrogen through
pipelines needs hydrogen-specific infrastructure such as hydrogen pipelines and compressors,
which can become an economic barrier. Thus, the idea of blending hydrogen into existing natural
gas pipelines arises as a potential alternative for transporting hydrogen economically by using
existing natural gas grids. However, there are several potential issues that must be considered when
blending hydrogen into natural gas pipelines. Hydrogen has different physical and chemical
properties from natural gas, including a smaller size and lighter weight, which require higher
operating pressures to deliver the same amount of energy as natural gas. Additionally, hydrogen's
small molecular size and lower ignition energy make it more likely to permeate through pipeline
materials and seals, leading to degradation, and its wider flammability limits make it a safety
hazard when leaks occur.
In this study, we investigate these potential issues through simulation and technical surveys.
We develop a gas hydraulic model to simulate the physical characteristics of a transmission and a
distribution pipeline. This model is used throughout the study to visualize the potential impacts of
switching from natural gas to hydrogen, and to investigate potential problems and solutions.
Furthermore, we develop a Real-Time Transient Model (RTTM) to address the compatibility of
current computational pipeline monitoring (CPM) based leak detection methods with blended
hydrogen. Finally, we suggest the optimal hydrogen concentration for this model, and investigate
the amount of carbon reduction that could be achieved, while considering the energy needs of the
system.

Keywords: Hydrogen Blending, Pipeline Safety, Leak Detection, Computational Pipeline


Monitoring
1
Nomenclature
𝐴 Cross-sectional area of pipe 𝜂𝑐 Compressor efficiency
𝛼 Proportional constant (HEE Index) 𝜌 Fluid density
𝑎𝑠 Speed of sound 𝑃 Pressure
𝐶𝑑 Discharge coefficient P1 Upstream pressure
𝐶𝑚 Carbon content, mixture P2 Downstream pressure
𝐶𝑟 Compression ratio Pb Atmospheric pressure
𝐶𝑠 Carbon content, hydrocarbon 𝑃𝑑 Discharge pressure
D Internal diameter 𝑃𝑠 Suction pressure
𝐷𝑓 Design factor 𝑃𝑜𝑤𝑒𝑟 Compression power
𝐷𝑜 External Diameter Q Volumetric flow rate
𝐸𝑓 Pipe seam joint factor 𝑄𝑖 Original flow rate
𝐸𝐶𝑂2 Equivalent CO2 emission 𝑄𝑟 Desired flow rate
𝑒𝑑 Absolute pipe roughness 𝑅𝑒 Reynolds number
𝑓 Darcy friction factor 𝑆 Minimum yield strength
FC Volume of fuel consumed 𝑆𝑔 Specific gravity of the fluid
G Gas gravity 𝑆ℎ Hoop stress
𝑔 Gravitational acceleration 𝑇 Temperature
ℎ Elevation 𝑇𝑓 Temperature derating factor
𝐻 Hydraulic head 𝑡 Time
∆𝐻𝑐 Compressor head 𝑡𝑝 Wall thickness
𝐻𝑓 Material performance factor 𝑇1 Suction temperature
𝐽𝑡 Joule-Thompson coefficient Tb Atmospheric temperature
L Length of pipe section Tf Average fluid temperature
𝐿𝐻𝑉𝐶𝐻4 Methane’s lower heating value 𝑉 Flow velocity
𝐿𝐻𝑉𝐵 Blend’s lower heating value 𝑊𝑡%𝑖 Weight percent of component
𝑀𝐴𝑂𝑃 Maximum allow. operating pressure 𝛾 Ratio of specific heats of gas
𝑀𝑓 Mass flow rate 𝛾𝑤 Specific weight
𝑀𝑤 Molecular weight Z Gas compressibility factor
𝑛 Material decaying exponent 𝑍1 Compressibility of gas (suction)
𝜂𝑎 Compressor adiabatic efficiency 𝑍2 Compressibility of gas (discharge)

2
1. Introduction
Fossil fuels, which are a primary energy source [1], emit greenhouse gases (GHG) during
the combustion process for thermal energy. This plays a significant role in causing climate change.
On the other hand, hydrogen is considered an alternative energy source because it produces zero
carbon emissions from combustion and electrochemical processes and has a high gravimetric
energy density [2]. This makes hydrogen a potential fuel for mitigating climate change by reducing
GHG emissions and providing effective energy. In addition, hydrogen can be produced through
electrolysis (Power to Gas: P2G) [3], where electricity decomposes water into hydrogen and
oxygen. This process can also be reversed to generate electricity through the reaction between
hydrogen and oxygen with fuel cells. Alternatively, hydrogen can be converted to electrical energy
through combustion (Gas to Wire: GTW) [4]. These diverse benefits of hydrogen usage can
significantly reduce carbon emissions and control energy balance by using surplus energy from
power generators or renewable energy production to produce hydrogen, which can then be
distributed as an energy carrier to meet high energy demands. To make sustainable hydrogen
distribution a reality in daily life, large-scale hydrogen production is necessary.
Transporting hydrogen can be an economic concern for appropriate distribution [5,6].
Trucks and ships can be used to transport hydrogen inside pressurized vessels, in either its
compressed gas (GH2) or liquid (LH2) form. Compressed gas requires relatively high pressures
(180 bar or higher) [7] and can require three to four times more infrastructure to be built [8]. Liquid
hydrogen is effective for increasing density by a factor of 800 [9], but it requires an energy-
intensive liquefaction process and special cryogenic insulated tanks to maintain the necessary low
temperatures. Alternatively, compressed liquid ammonia can be used to transport bulk hydrogen
to remote areas through trucks or ships.
The economic supply of energy for hydrogen conversion is dependent on the availability
of local resources and the method of transportation. Pipelines are considered the most cost-
effective way to transport large volumes of hydrogen, with estimates ranging from $0.05 to $3 per
ton depending on the distance [5]. Trucks and ships are also viable alternatives for hydrogen
transportation, but the low density of hydrogen requires it to be pressurized into a compressed gas
(around 18 MPa) [6] or liquified for high density energy, which increases costs due to the necessary
infrastructure and processes for compression and liquefaction.
While pipelines are the most cost-effective way in the long-term, they require specialized

3
infrastructure, such as hydrogen pipelines and compressors, which can pose an economic barrier.
One of the promising solutions to this barrier is blending hydrogen with natural gas into existing
pipeline networks. However, the different properties of blended gas can lead to operational,
material, and safety issues that require modification of the existing system [10,11]. The schematic
of potential blended hydrogen pathways is shown in Figure 1.

Figure 1. Blended Hydrogen Pathways.

There have been numerous studies on the transportation of hydrogen in pipelines, ranging
from the Fort Saskatchewan Hydrogen Blending project, which uses a 5% blended H2
concentration [6], to H21, which uses 100% hydrogen concentration [12]. Table 1 illustrates a non-
exhaustive list of blended hydrogen transportation projects.

Table 1. List of Major Projects regarding Blended Hydrogen Transportation.


Blended
No Project Country Objectives References
H2 [%]
Blending Hydrogen into natural
1 HyDeploy UK 20% [11,13]
gas pipelines
Pilot program for blending
Fort Saskatchewan hydrogen into natural gas
2 Hydrogen Blending Canada pipelines that serve residential 5% [6]
Project (ATCO) and commercial buildings in
Fort Saskatchewan.

4
Investigation of 100% Hydrogen
3 H21 UK usage with existing natural gas 100% [12]
system
Technical evaluation for
4 Hyblend US blending hydrogen in natural - [14]
gas pipelines
Blending Hydrogen into Gas
5 GRHYD France 20% [14]
Distribution Lines
Blending Hydrogen into Gas
6 Snam Italy 10% [14]
Transmission Gas Lines
Green Hydrogen Blending into
7 HyP SA Australia 5% [15]
Gas Distribution Lines
Blending hydrogen into gas
Enbridge and Cummins
distribution lines in Markham,
8 Hydrogen-blending Canada < 2% [16]
Ontario and performing a
Project
routing study
Comprehensive assessment of
using hydrogen as a fuel in
9 Hy4Heat UK 100% [14]
residential and commercial
buildings
Blending hydrogen into gas
Hydrogen injection in the
10 Denmark distribution in Varde, Denmark, 15% [17]
gas grid
and safety evaluation
Injecting renewable hydrogen
11 Cleangas Turkey Turkey into natural gas and gas mixture 20% [17]
evaluation
EN-H2
Carbon neutralization with the
12 (Portugal National Portugal 15% [18]
hydrogen economy
Hydrogen Strategy)

Many studies and projects have examined the general issues surrounding the transportation
of blended hydrogen through existing gas pipelines and have addressed potential risks from
operation and failure events. However, there is a need for more in-depth research on operation,
material, and safety matters.
A key question in the transportation of blended hydrogen is whether existing natural gas
pipelines can be used, considering the limits of hydrogen concentration and corresponding blended
gas changes due to operating pressure. These factors can have significant impacts on compression,
friction, temperature, and energy supply [19]. Additionally, typical natural gas pipeline materials
may be susceptible to hydrogen embrittlement and require a reduction in operating pressure. The
low density of hydrogen can also lead to different gas dynamics in the case of leak events, making
conventional leak detection systems potentially incompatible with blended hydrogen pipelines.
Moreover, hydrogen leakage into a confined area has the potential for significant damage due to

5
the wide detonation limits of hydrogen. Ultimately, the aim of blending hydrogen is to reduce
greenhouse gas (GHG) emissions.
The goal of this study is to examine the potential advantages and challenges of using blended
hydrogen gas in existing natural gas systems. To do this, we conduct a thorough review of the
literature on the effects of hydrogen on pipeline materials, safety considerations such as leak
detection, potential consequences of leaks, and GHG emissions. Additionally, we investigate a gas
hydraulic model to identify the optimal hydrogen blends for both transmission and distribution
pipelines as part of a case study, taking into consideration the impacts on the pipeline, compressor
selection, and other design variables. We also develop a Real-Time Transient Model (RTTM) that
incorporates the simulation parameters and demonstrates how pipelines with different mixtures of
hydrogen and natural gas would behave. Finally, we model the expected GHG emissions with the
pipeline case study, considering the impact not only of hydrogen combustion, but also of its
production. Through this investigation we further elucidate the feasibility of blended hydrogen gas
in everyday use and the potential for GHG emission reductions.

2. Blending Hydrogen and Natural Gas


Blending hydrogen with natural gas in existing pipeline systems is the most economical way
to transport large volumes of hydrogen over long distance without the need for new infrastructure.
However, the small size of the hydrogen molecule and its different physical properties, such as
lower volumetric density and viscosity, can lead to unique behavior when mixed with natural gas
[20,21]. This can pose potential safety risks for pipelines that were designed and built to transport
natural gas. To maintain energy balance and depending on hydrogen concentration, the blended
gas mixture may need to be transported at higher flow rates than natural gas. This may result in
higher operating pressures and potentially exceed the design limits of compressors and pipelines,
which were originally designed for natural gas transmission and distribution. Therefore, it is
important to consider design changes to ensure the safe transportation of blended hydrogen through
existing pipeline systems and identify any potential risks and operational issues related to hydrogen
concentration.

2.1 Hydrogen Blending Methods


It is important to ensure that blended gas is in a homogeneous state, with uniform behavior

6
along the entire length of the pipeline [22]. If the two gases have significantly different densities,
they may stratify, leading to different flow behavior and leak characteristics. This can result in
uneven energy distribution and operational issues with the pipeline. A typical system for blending
hydrogen into natural gas grids is shown in Figure 2 and includes an electrolyzer, hydrogen buffer
storage, and a blending unit.

Figure 2. The Schematic of a Typical Blending Hydrogen into Natural Gas Grid.

There are several hydrogen blending systems in operation. Table 2 lists some of these
projects. Australia's HyP SA project [15] uses hydrogen purification equipment between the
electrolyzer and the hydrogen buffer tank to remove produced impurities. ATCO's CEIH project
[23] in Australia blends up to 25% hydrogen using a flow control system with gas chromatography
(GC) and a PID controller. The HyDeploy project [11] uses a mixing loop to blend hydrogen with
natural gas to produce desired hydrogen concentrations in the blended gas before injecting it into
natural gas pipelines. The blending unit measures the components of the blended gas with a gas
control system to ensure that the two gases are fully mixed.
The Jupiter 1000 project [24], which was commissioned in 2018 in France, generates
hydrogen using two electrolyzers with a combined capacity of 1.0 MW and methane through a
methanation process that converts captured CO2. These gases are then blended to the desired
concentration and injected into natural gas pipeline for distribution.

7
Table 2. Examples of Hydrogen Blending Systems. [11,15,23–25]
Electrolyser Hydrogen
Project Country Network Others
Capacity Blend %
HyP SA Australia Distribution 1.2 MW 5% Uses green hydrogen

ATCO-CEIH Australia Distribution 0.15 MW 5-25% GC and PID Control


Measures Wobbe
HyDeploy UK Distribution 0.5 MW 20%
index
Injects hydrogen into
Jupiter 1000 France Transmission 1.0 MW 6%
transmission lines

Finally, lab-scale equipment has also been developed to blend hydrogen with natural gas
at lower flow rates by rotating helical fins inside a pipe [22]. All of these hydrogen blending
methods are used to produce fully developed gas mixtures before they are transported through
pipelines to prevent gas stratification. When blended gas is transported under turbulent flow
conditions, hydrogen and natural gas do not stratify in the pipeline, as investigated in the HyDeploy
project [11]. For laminar or stagnant flows, possible scenarios for distribution lines, gas
stratification can be a potential risk due to density differences between the mixture components.

2.2 Transmission and Distribution Pipeline Simulation Properties


In order to examine the potential conversion of a natural gas pipeline to a blended hydrogen
pipeline, this study uses a simulation to analyze various design parameters. The focus of the case
study are two sample pipelines: a 322 km transmission line and a 10 km distribution line, as shown
in Figure 3. Transmission lines generally operate at higher pressures and transport larger volumes
over long distances [21], while distribution lines tend to operate at lower pressures and cover
shorter distances.

Figure 3. Illustration of the simulation pipelines - Transmission and distribution pipelines.

The design specifications for the simulated pipelines are largely based on Case Study 6 in
"Transmission Pipelines Calculations and Simulations Manual" [27], but with a different pipe
8
material and weld type. This choice was made in order to make the pipeline selection as realistic
as possible. The full specifications can be found in Table 3.

Table 3. Design Specification for the Pipelines used in the Simulation.


Contents Transmission Line Distribution Line

Fluid H2: 0 ~ 100%, CH4: 100% ~ 0%


Inlet Temperature 20 °C
Location Underground (Fully Insulated)
Material API 5L X70 ASTM A53 Gr. B
Tensile Strength (MPa) 565.37 330.95
Diameter (I.D, mm) 609.6 (NPS 24) 102.2604 (NPS 4)
Weld Type ERW Seamless
Wall Thickness (mm) 12.7 6.0198
Surface Roughness (mm) 0.01778
0.41 (60 psig)
Delivered Pressure (MPa) 3.45 (500 psig)
(Before Service Regulator)
14,158,423
Operating Flow (m3/day) 15,000
(500MMSCFD)
Length (km) 322 10

In transmission pipelines, gas is usually transported using centrifugal compressors [28,29].


The number and location of compressors are chosen to ensure that the system is hydraulically
stable and can safely deliver the gas flow. It is important to design the gas compressors with
optimal compression ratios and head limits for safe transportation.
For the purpose of this simulation, we have assumed that there is only one compressor in
the transmission line, but no compressor in the distribution line. The reason for this decision is that,
as the simulation progresses, it becomes clear that the transmission line is often the limiting factor
and therefore the focus of the study. The compressor for station 1 was defined based on a typical
transmission compressor, and the details are shown in Table 4.

9
Table 4. Design Specification for the Compressor used in the Simulation.
Contents Transmission Line Distribution Line

Fluid H2: 0 ~ 100%, CH4: 100% ~ 0%


Suction Pressure (MPa) 3.45 (500 psig) N/A1*
Compressor Adiabatic
0.85 N/A1*
Efficiency
Maximum Head Per Stage (m) 2,438 (8,000 ft) N/A1*
Suction Temperature 15 °C N/A1*
Discharge Temperature 20 °C N/A1*
Discharge Pressure (MPa) 8.34 (1,210 psig) N/A1*
1
* No compressor in the distribution line (Pressure reduction station only).

Natural gas typically consists of more than 95% methane, with other gases making up
smaller fractions [26], as shown in Table 5. For this study, we assume that natural gas is completely
composed of methane, as this assumption is not expected to significantly affect the interpretation
of the results.
Hydrogen and methane have different physical and chemical properties. These properties
include molecular weight, density, viscosity, and specific gravity. Hydrogen has a molecular
weight that is 8 times lower than methane, which results in significantly lower values for most
physical properties except viscosity; the viscosity of hydrogen is approximately 30% lower than
methane at room temperature (25°C).
Hydrogen contains 2.4 times more energy per unit mass than methane. However, the lower
heating value (LHV) of hydrogen per unit volume is three times lower than methane due to its low
density at standard conditions. This means that the energy content of the blended gas may be
reduced with higher concentrations of hydrogen. From the safety perspective, higher
concentrations of hydrogen tend to increase the risk of fire and explosion. Compared to methane,
pure hydrogen has a much broader flammability range (5.3 times) and detonation limit range (7.1
times). It also has a significantly lower ignition energy (14.5 times lower), which makes it more
prone to ignition and increases the risk of fire.
The low density of hydrogen also leads to higher diffusivity with other materials, which can
result in more severe flammable and detonable events in confined spaces. Table 6 summarizes the

10
properties of each gas at 20°C and 101.35 kPa in terms of physical properties, energy content and
risk.

Table 5. Typical Natural Gas Composition. [26]


Methane Ethane Nitrogen Carbon Dioxides Propane N-Butane Isobutane
Constituent Others
(CH4) (C2H6) (N) (CO2) (C3H8) (C4H10) (C4H10)

Volume (%) 95.3 2.16 1.86 0.44 0.19 0.02 0.02 0.01

Table 6. Gas Properties – Comparison between Hydrogen and Methane (at CNTP).

Properties Hydrogen Methane Unit Reference

Molecular Weight 2.016 16.043 kg/kmol [30]

Density (Gas) 0.08 0.65 kg/m3 [31]

Density (Liquid) 71 430-470 kg/m3 [31]

Specific Gravity 0.0696 0.555 - [12]

Viscosity (at 25°C) 0.89 1.11 10-5 Pas [20]

Diffusion Coefficient in Air 0.61 0.16 cm2/s [32]

Energy Density 120 50 MJ/kg [31]

Lower Heating Value (LHV) 10.2 34 MJ/m3 [21]

Higher Heating Value (HHV) 12.5 37.8 MJ/m3 [21]

Flammability Limits (in air) 4-75 5-15 Vol % [31]

Detonation Limits (in air) 18.3-59 6.3-14 Vol. % [20]


Minimum Spark Ignition
0.02 0.29 MJ [30]
Energy
Auto Ignition Temperature 858 810 K [30]

Laminar flame Velocity 2.1 0.4 m/s [21]

Specific Heat Capacity 14.86 2.22 J/(gK) [33]

Specific Heat Ratio 1.383 1.308 - [33]

Flame Temperature in Air 2318 2148 K [33]

11
Solubility in Water 0.0016 0.025 kg/m3 [33]

2.3 Energy Transmission


Energy transmission is a key factor to consider when converting a natural gas pipeline to
use blended hydrogen. The amount of transmitted energy can be calculated by multiplying the
standardized operating flow rate by the lower heating value (LHV) of the blend. When using pure
hydrogen, the energy transmission is approximately 3.3 times lower than when using methane. In
other words, the lower the concentration of hydrogen in the blend, the higher the energy
transmission will be, for the same flow rate and pressure.
This is illustrated in Figure 4 for the simulated (a) transmission and (b) distribution
pipelines. For example, the transmission pipeline with natural gas would typically transmit
4.80x108 MJ/day, but this would drop to 1.44x108 MJ/day with pure hydrogen in the pipeline. The
figure shows the full range of variation for different hydrogen concentrations.

(a) Transmission Pipe at 14,158,423 m3/day (b) Distribution Pipe at 15,000 m3/day
Figure 4. Energy Transmission by Hydrogen Concentration in the Gas Mixture.

Thus, if it is desired to maintain the energy transmission the same in a pipeline conversion,
the flow rate and/or operating pressure must be increased accordingly. These changes may not be
trivial. Increasing the operating pressure may be constrained by the pipe material used, or it may

12
be limited by the number of compressors that can be installed.

2.4 Gas Hydraulics


The flow of gas in pipelines depends on various factors, such as pressure, temperature, and
the physical characteristics of the pipeline, such as length and diameter. The physical relationships
that govern this flow in a steady and isothermal state can be described using the general gas flow
equation [34] shown in Eq. (1).

2
𝑄 ⋅ 𝑃𝑏
𝑃12 = 𝑃22 + ( ) 𝐺 ⋅ 𝑇𝑓 ⋅ 𝐿 ⋅ 𝑍 ⋅ 𝑓 (1)
1.1494 ⋅ 10−3 𝑇𝑏 ⋅ 𝐷2.5
where Q is the volumetric flow rate (m3/day), 𝑓 is the Darcy friction factor (dimensionless), Pb is
atmospheric pressure (kPa), 𝑇𝑏 is atmospheric temperature (K), P1 and P2 are the absolute pressure
(kPa) at upstream and downstream locations, G is gas gravity (m/s2), 𝑇𝑓 is average gas flowing
temperature (K), L is the length of the pipe (km), Z is gas compressibility factor at the flowing
temperature (dimensionless), and D is the inner diameter of the pipe (mm).
This section analyzes the impact of hydrogen concentration on various properties,
including friction, temperature, flow rate changes, and pressure, and provides guidelines for
determining the optimal amount of hydrogen to blend in the pipelines investigated in the case study.
All of the parameters are visualized in relation to hydrogen concentration and energy transmission.

2.4.1 Flow Rate

The flow rate depends on the desired energy transmission and hydrogen concentration in
the blend. As discussed in Sub-Section 2.3, the energy transmitted can be calculated by multiplying
the LHV of the gas mixture by its flow rate. The required flow rate to maintain a certain level of
energy transmission can then be calculated using Eq. (2).

𝑄𝑖 ⋅ 𝐿𝐻𝑉𝐶𝐻4
𝑄𝑟 = , (𝑖 = 1, 2) (2)
𝐿𝐻𝑉𝐵
where Qr is the desired volumetric flow rate, Qi is the original volumetric flow rate (defined in
Table 3), I denotes the type of pipeline (transmission or distribution), and 𝐿𝐻𝑉𝐶𝐻4 and LHVB are
the lower heating values for methane and the blended gas (methane and hydrogen) respectively.

13
Figure 5 shows the variation in flow rate for (a) the transmission and (b) the distribution
pipelines as a function of hydrogen concentration and energy transmission. The energy
transmission is plotted on a scale from 30% to 100%, with values corresponding to those shown
in Figure 4 (i.e., 30% corresponds to 1.44x108 MJ/day, and 100% to 4.80x108 MJ/day). As
expected by intuition, a significant increase in flow rate would be needed to maintain high energy
transmission and hydrogen concentration. This is relevant for all of the following factors.

(a) Transmission Pipeline (b) Distribution Pipeline


Figure 5. Flow Rate with Hydrogen Concentration and Energy Transmission.

2.4.2 Friction Factor

In laminar flow, fluid friction is only dependent on the Reynolds number, as shown in Eq.
(3). For turbulent flow, it can be described using the Colebrook-White equation [34] shown in Eq.
(4). This equation considers pipeline roughness (assumed to be constant), pipeline diameter, and
Reynolds number.

64
𝑓= (3)
𝑅𝑒
1 𝑒𝑑 2.51
= −2𝐿𝑜𝑔10 ( + ) (4)
√𝑓 3.7𝐷 𝑅𝑒√𝑓

where ed is absolute pipe roughness (mm) and Re is the Reynolds number (dimensionless). The
friction factor (𝑓) is important because it is used to determine the upstream and downstream

14
pressures according to Eq. (1). Additionally, a high friction factor is generally undesirable as it can
lead to higher pressure drops and/or lower flow rates, potentially requiring modifications to the
pipeline design.
Figure 6 illustrates variation of the friction factor for (a) the transmission and (b) the
distribution pipelines as a function of hydrogen concentration and energy transmission. It is clear
that to maintain the same level of energy transmission with a higher hydrogen concentration, the
friction factor increases significantly, indicating a higher Reynolds number and a more turbulent
flow.

(a) Transmission Pipeline (b) Distribution Pipeline


Figure 6. Friction Factor variation with Hydrogen Concentration and Energy Transmission.

2.4.3 Temperature

Fluid transport through pipelines can result in temperature variations due to two main
sources. The first is the heat transfer between the inside of the pipeline and its surroundings. The
second is the Joule-Thompson effect, which is a change in temperature that occurs when gas
expands due to pressure differences between the inlet and outlet [35]. This effect is described by
Eq. (5).

𝜕𝑇
𝐽𝑡 = ( ) (5)
𝜕𝑃 𝐻
where 𝐽𝑡 is the Joule-Thompson coefficient (°C/bar), T is the temperature (°C), and P is pressure

15
(bar). H denotes that the partial derivate is evaluated at constant Enthalpy. The Joule-Thompson
coefficient (𝐽𝑡 ) is a property of the gas being used as well as the temperature and pressure of the
gas prior to expansion. Every gas has an inversion point at which the Joule-Thompson coefficient
changes sign and the gas begins to either warm or cool. Most gases, including methane, have a
high inversion point temperature, meaning that a pressure drop has a cooling effect. Hydrogen,
however, is one of the few gases that has a very low inversion point, with an inversion point of
202 K at atmospheric pressure [35]. As a result, hydrogen generally experiences a warming effect
when subjected to a pressure drop.
In the simulation, the Joule-Thompson effect is accounted for by assuming constant values
for methane (-0.5°C/bar) and hydrogen (0.035°C/bar), obtained from the literature [36]. The
impact of this is shown in Figure 7 for both transmission and distribution pipelines. As expected,
the highest temperature occurs at higher concentrations of hydrogen and transmitted energy. It is
also noteworthy that the range of temperature variation in the transmission pipeline (from -20°C
at 100% methane / 30% energy to 30°C at 100% hydrogen / 100% energy) is much larger than that
in the distribution pipeline (from 19.6°C to 20.1°C in the same cases). This is due to the higher
volumetric flow rate in the transmission line.

(a) Transmission Pipeline (b) Distribution Pipeline


Figure 7. Temperature Variation due to Joule-Thompson Effect with Transmitted Energy and
Hydrogen Concentration.

16
2.4.4 Maximum Allowable Operating Pressure (MAOP)

The operating pressure is closely related to material stress. This is especially important
when converting existing methane pipelines for hydrogen transportation, as engineers must
consider the existing pipeline material. Standard design guidelines for pipelines typically state that
the governing stress is the hoop stress (circumferential direction), which can be described
mathematically using the following equation [34].

𝑃 ⋅ 𝐷𝑜
𝑆ℎ = (6)
2𝑡𝑝

where Sh is the pipe’s hoop stress (Pa), 𝑡𝑝 is the pipe wall thickness (mm), 𝐷𝑜 is the external
diameter (mm), and P is the internal operating pressure (Pa). This equation shows that the hoop
stress is a function of the operating pressure, pipeline diameter, and design thickness. This
demonstrates that the operating pressure directly affects the stress level in the pipe. The practical
relationship between these variables can be found in industrial standards [37,38] and typically also
includes design and safety factors. For hydrogen pipelines, the maximum allowable operating
pressure (𝑀𝐴𝑂𝑃) can be represented by the following equation [38].

2𝑆 ⋅ 𝑡𝑝
𝑀𝐴𝑂𝑃 = 𝐷𝑓 ⋅ 𝐸𝑓 ⋅ 𝑇𝑓 ⋅ 𝐻𝑓 (7)
𝐷𝑜
where 𝑀𝐴𝑂𝑃 is the maximum allowable operating pressure (Pa), S is the minimum yield strength
(Pa), 𝑡𝑝 is the wall thickness (mm), 𝐷𝑜 is the external diameter (mm), 𝐷𝑓 is the design factor
(dimensionless), 𝑇𝑓 is the temperature derating factor (dimensionless), 𝐸𝑓 is the pipe seam joint
factor (dimensionless), and Hf is the material performance factor (dimensionless).
This equation can be used to determine the operating pressure limit for hydrogen use in the
current gas pipe design, which will naturally be lower than when the pipe is used for natural gas
service. With the pressure limit, the inlet pressure can be calculated with the general flow equation
(Eq. (1)) for different hydrogen concentrations and transmitted energy. This is shown graphically
in Figure 8.
A black shaded square area is included in (a) the transmission pipeline to indicate the
MAOP (maximum allowable operating pressure) for this pipeline, indicating that it would not be

17
possible to maintain the same transmission energy with pure hydrogen because it would exceed
the MAOP and pose a structural risk. In contrast, the MAOP is not reached in (b) the distribution
pipeline, so it would be safe to operate within the entire range of hydrogen concentration and
transmitted energy.

(a) Transmission Pipeline (b) Distribution Pipeline


Figure 8. Inlet Pressure Variation with Hydrogen Concentration and Energy Transmission.

2.4.5 Compressor Selection

In the previous sections, we examined the potential impacts of converting a natural gas
pipeline to hydrogen in terms of operating pressure and flow rate, but we did not address how these
parameters would be achieved. This section focuses on the selection of compressors that would be
compatible with such parameters, and whether or not the scenarios would be feasible.
The first parameter to be considered is the compression ratio, which is basically a measure
of how much the compressor can increase the suction pressure. It can be calculated with the
following equation.

𝑃𝑑
𝐶𝑟 = (8)
𝑃𝑠
where Cr is the compression ratio (dimensionless), and Pd and Ps are the discharge and suction
pressures (Pa). The compression ratio assumed for the simulation is 1.5, a common value for
centrifugal compressors. This limit is used to determine the number of compressors required [34].
For example, if a compression ratio of 2.8 is required for a particular scenario, since this is higher

18
than 1.5, two compressors are necessary. With this compression ratio set as the limit, Figure 9
illustrates the compression ratio and number of compressors necessary to achieve different
hydrogen concentrations and transmitted energies. The black shaded square represents regions that,
if above, would require more than one compressor.

Figure 9. Compression Ratio Variation with Hydrogen Concentration and Transmitted Energy.

The other two parameters that need to be considered are compressor head (∆𝐻𝑐 ) and
compressor power (𝑃𝑜𝑤𝑒𝑟), two interconnected parameters with the flow rate that help determine
the number of compressor stages required by the system. The general expression for the
compressor head are found in Eq. (9) and Eq. (10) [34].

𝛾−1
𝛾 𝑍1 + 𝑍2 1 𝑃2 𝛾
𝑃𝑜𝑤𝑒𝑟 = 4.0639 ( ) 𝑄 ⋅ 𝑇1 ( ) ( ) [( ) − 1] (9)
𝛾−1 2 η𝑎 𝑃1

𝑀𝑓 × ∆𝐻𝑐
𝑃𝑜𝑤𝑒𝑟 = (10)
𝜂𝑐
where Power is compression power (kW), 𝛾 is the ratio of specific heats of the gas (dimensionless),
Q is the volumetric flow rate (Mm3/day), T1 is the suction temperature of the gas (K), P1 and P2
are suction and discharge pressures of the gas (kPa), Z1 and Z2 is the compressibility of gas at
suction and discharge conditions (dimensionless), η𝑎 is the compressor adiabatic efficiency
(dimensionless), Mf is the mass flow rate of gas (kg/s), ∆𝐻𝑐 is the compressor head (m), and 𝜂𝑐 is

19
the compressor efficiency (dimensionless). The compressor efficiency (𝜂 a) is the compressor
adiabatic efficiency.
Using these equations, the compressor power and head are illustrated, respectively, on
Figure 10, (a) and (b). As with the previous image, the black shaded square represents the limit
(maximum head per stage). As expected, a higher hydrogen concentration and transmitted energies
requires a significant increase in compressor power. It is interesting to see how many more
compressor stages would be required to transport pure hydrogen at high energy deliveries. To
achieve 100% transmitted energy with pure hydrogen, a total of 11 compressor stages would be
required.

(a) Compressor Power (b) Compressor Head and Compressor Stages


Figure 10. Compressor Power, Head and Number of Compressor Stages with Hydrogen
Concentration and Transmitted Energy.

2.5 Optimal Hydrogen Blend for Pipeline Transportation


In conclusion, finding the optimal blend of hydrogen and natural gas involves tradeoffs, as
increasing the percentage of hydrogen may require additional equipment and modifications to the
pipeline. For example, transmitting 100% energy with pure hydrogen would require a large number
of compressors and pipeline modifications, while transmitting 65% energy with a 60% hydrogen
blend would only require the addition of a single compressor. Alternatively, transmitting 51% of
the energy with a 30% hydrogen blend would not require any modifications at all. These examples,
as well as a wider range of options described in Table 7, demonstrate the various tradeoffs involved
in determining the optimal blend. The best choice will depend on the specific project goals and

20
budget constraints, and a careful evaluation of the available options is necessary to determine the
most appropriate blend for a given situation.

21
Table 7. Optimal Hydrogen Concentration without System Modification (Transmission Line).

Hydrogen Concentration without Modification (%)


Gas Pressure (MPa) Compression Ratio Compressor Head (m) No. of
Energy
Contents Compressor
Content Simulation Simulation Simulation Modification
(H2 – CH4) Limit [45] Limit [40] Limit s
Results Results Results required
0 – 100% 12.27 2.20 1915.95 2 Pipe, Compressor
20 – 80% 13.08 2.35 2807.14 2 Pipe, Compressor
40 – 60% 13.93 2.50 4171.38 2 Pipe, Compressor
100% 11.23 1.5 2438
60 – 40% 14.80 2.65 6511.26 3 Pipe, Compressor
80 – 20% 15.51 2.78 11372.60 5 Pipe, Compressor
100 – 0% 14.83 2.66 26730.60 11 Pipe, Compressor
0 – 100% 10.92 1.96 1617.67 1 Compressor
20 – 80% 11.59 2.08 2384.00 1 Pipe, Compressor
40 – 60% 12.30 2.21 3560.97 2 Pipe, Compressor
86% 11.23 1.5 2438
60 – 40% 13.04 2.34 5582.66 3 Pipe, Compressor
80 – 20% 13.63 2.44 9775.49 5 Pipe, Compressor
100 – 0% 13.06 2.34 22846.72 10 Pipe, Compressor
0 – 100% 9.01 1.62 1140.21 2 Compressor
20 – 80% 9.48 1.71 1700.39 2 Compressor
40 – 60% 9.98 1.79 2567.37 2 Compressor
65% 11.23 1.5 2438
60 – 40% 10.50 1.89 4062.70 2 Compressor
80 – 20% 10.92 1.96 7155.38 3 Compressor
100 – 0% 10.51 1.89 16540.60 7 Compressor
0 – 100% 8.42 1.42 810.52 1 -
20 – 80% 8.82 1.48 1221.56 1 -
40 – 60% 9.25 1.54 1862.81 2 Compressor
51% 11.23 1.5 2438
60 – 40% 9.69 1.61 2974.17 2 Compressor
80 – 20% 10.06 1.66 5269.17 3 Compressor
100 – 0% 9.709 1.61 12064.05 5 Compressor
0 – 100% 6.43 1.16 344.12 1 -
20 – 80% 6.57 1.19 528.94 1 -
40 – 60% 6.73 1.22 822.41 1 -
30% 11.23 1.5 2438
60 – 40% 6.90 1.25 1337.38 1 -
80 – 20% 7.04 1.27 2400.37 1 -
100 – 0% 6.90 1.25 5395.90 3 Compressor

22
3. Hydrogen influence on Pipeline Materials
One major concern with using hydrogen in pipelines is its ability to permeate through
materials. As the smallest and lightest element (see Table 6), hydrogen can easily diffuse into the
lattice structure of materials, causing them to become weaker and more prone to cracking under
stress [38]. This process, known as hydrogen embrittlement (HE), can be a significant issue for
pipelines and lead to unexpected failures. This section discusses the consequences of this process
and potential solutions.

3.1 Hydrogen Embrittlement


Hydrogen Embrittlement can reduce mechanical properties such as tensile ductility,
fracture toughness, and fatigue resistance, making them more prone to loss of ductility and
cracking [38]. A report from the US Department of Energy [39] indicates that the majority of
natural gas networks, including transmission and distribution lines, are made of carbon steel
materials (i.e., 99.7% and 50.4%, respectively). As such, when considering the use of existing
natural gas networks for hydrogen transportation, it is crucial to determine whether the existing
metallic materials are compatible with hydrogen gas service.
Carbon steels, including ASTM A 106 Grade B, ASTM A 53 Grade B, and API 5L Grades
X42 and X52, have been shown to be suitable for use in hydrogen pipelines with pressures up to
14 MPa, according to ASME B31.12 [38]. Table 8 summarizes the design specifications of
hydrogen pipelines currently in operation worldwide. These pipelines are usually made of carbon
steels and operate at a maximum pressure of up to 13.1 MPa. This helps contextualize and provide
a practical perspective on the materials and pressures currently being used in hydrogen pipelines.

Table 8. Operating Pressure for Hydrogen Pipelines Worldwide.


Operating
Length Diameter
Location Pressure Material Operator Reference
(Km) (mm)
(MPa)
Air
Texas, US 560 150-305 2.4-13.1 Steel Pipe [21]
Products
Air [21]
Westliche, US 367 203-254 2.4-13.1 Steel Pipe
Products
Air [21]
Louisiana - 150,203,305 2.4-13.1 Steel Pipe
Products
Air [21]
Los Angeles 19 150,254,305 2.4-13.1 Steel Pipe
Products

23
API 5L Air [21]
Texas, US 105 203 2.4
Grade B Liquide
API 5L Air [21]
Texas, US 55 36 5.1
X60 Liquide
Seamless
Air
Netherlands 879 304.8 6.5~10 Carbon [40]
Liquide
Steel
Canada 3.7 273 3.8 Gr.290 AGEC [40]
Carbon ICI [40]
UK 15 - 30
Steel Billingham
Germany API 5L [21]
100 - 2-5 Linde
(Leuna) L290
France, API 5L Air [21]
1040 100 9.7
Belgium X52 Liquide

The HEE (Hydrogen Environment Embrittlement) Index is a metric that can be used to
evaluate the potential for hydrogen embrittlement in a material. It compares the Notched Tensile
Strength (NTS) of a material when subjected to hydrogen to when it is subjected to air or helium.
The HEE Index can be calculated using the equation:

HEE Index = NTS Ratio = NTS in Hydrogen/NTS in Air or Helium (11)

The HEE index is a ratio that ranges from 0 to 1, with higher values indicating a lower
potential for hydrogen embrittlement. Materials with an HEE index as close to 1 as possible are
typically desired. According to NASA’s report [41], Table 9 categorizes different levels of severity
based on the HEE index. Detailed test procedures for obtaining the HEE index can be found in the
literature [42].

Table 9. The Categorization of HE Severity for Industrial Materials. [41]


HE HEE Index
Description
Severity (NTS Ratio)
Negligible 1.0 – 0.97 Materials can be utilized in the specified hydrogen
pressure and temperature range with fracture and crack
Small 0.96 – 0.90 growth analysis in hydrogen
Limited application with fracture and crack growth
High 0.89 – 0.70
analysis in hydrogen
Severe 0.69 – 0.50 Not recommended for usage

24
Extreme 0.49 – 0.0

Table 10 summarizes the severity of the HEE Index and the compatibility of common
industrial materials with hydrogen. It also includes the hydrogen pressure at which the HEE Index
was obtained, which can be either 6.9 MPa or 68.9 MPa [38,41]. It is noteworthy that the results
show that ferritic steels commonly used in current gas pipelines have a range of HEE severity from
high to low when used at pressures up to 6.9 MPa at room temperature. Ductile materials such as
stainless steel (304 and 316) and copper-based materials have relatively low HEE severity, even
at high pressures (68.9 MPa). Additionally, these materials have good resistance to hydrogen
embrittlement in both gas and liquid states.

Table 10. HE Index for commercialized materials.


Material Compatibility
Lab-scale HE Evaluation
with Hydrogen [9]
Material HEE Reference
Hydrogen HE Index
Gas Liquid
Pressure Severity (NTS
ratio*2)
A106-Gr. B 6.9 MPa High - Acceptable Not Acceptable [38,41]

A516 6.9 MPa High 0.83 Acceptable Not Acceptable [38,41]

Ferritic API 5L
6.9 MPa High - Acceptable Not Acceptable [38]
Steels X42
(Carbon API 5L
6.9 MPa High 0.86 Acceptable Not Acceptable [38,41]
steels and X52
low alloy API 5L
steels) 6.9 MPa Small 0.92 Acceptable Not Acceptable [38,41]
X60
API 5L
6.9 MPa Small 0.94 Acceptable Not Acceptable [38,41]
X65
API 5L
6.9 MPa Small 0.90 Acceptable Not Acceptable [38,41]
X70
304L 68.9 MPa High 0.87 Acceptable Acceptable [38,41]
Austenitic
Steels
316 68.9 MPa Negligible 1 Acceptable Acceptable [38,41]
Copper
68.9 MPa Negligible 1 Acceptable Acceptable [38,41]
Copper (OFHC*1)
Based 70-30
68.9 MPa Negligible - Acceptable Acceptable [38]
Brass
*1: OFHC: Oxygen-free high thermal conductivity
*2: NTS Ratio is calculated in hydrogen and helium environments

25
The HEE index can also be correlated with the operating pressure through a proportionality
constant and a material-dependent exponential, as described in Eq. (12) [41]. This relationship
reinforces that higher pressures increase the risk of hydrogen embrittlement.

𝐻𝐸𝐸 𝐼𝑛𝑑𝑒𝑥 = 𝛼(𝑃)−𝑛 (12)


where α is a proportional constant, P is the hydrogen pressure (MPa), and n is the material decaying
exponent (material dependent).
Temperature is another factor that strongly influences hydrogen embrittlement behavior.
In general, hydrogen embrittlement occurs at temperatures below 95°C. At high temperatures
(>200°C), hydrogen embrittlement is usually not a problem for most materials [43]. However,
High Temperature Hydrogen Attack (HTHA), also known as hydrogen attack, can be a problem.
This occurs when hydrogen dissociates into its atomic form and diffuses into steel, reacting with
the carbon to form methane. This can cause various problems, including fissuring and cracks in
the pipeline. While each material’s resistance to HTHA varies and depends on both temperature
and pressure, API 941 provides a comprehensive list of recommended practices and operational
limits for different steels [43].
In summary, hydrogen embrittlement is a complex problem with a complex mechanism.
However, the resistance of many materials to hydrogen embrittlement has already been catalogued,
and various standards provide best practices for proper material selection. Choosing a material
known to have low susceptibility to embrittlement and maintaining internal pressure within
recommended limits is the optimal solution. In cases where the engineer has no control over
material selection, such as when reusing a natural gas pipeline, internal coatings can be used as an
alternative solution.

3.2 Internal Coatings to Prevent Hydrogen Embrittlement


Internal coatings offer a potential solution for enabling the use of materials that are not
currently compatible with hydrogen in pipeline transportation, or for improving their compatibility.
These coatings function as a barrier by mechanically separating the material from the hydrogen
environment [44,45]. Many of these coatings are still being researched and can be broadly
classified into metal, ceramic, and polymer coatings [44]. Additionally, two-dimensional materials
such as graphene and Mxene have been identified as promising hydrogen barrier coating materials

26
due to their excellent hydrogen resistance and mechanical durability [46,47]. The benefits and
drawbacks of each coating type are summarized in Table 11.

Table 11. Types of Barrier Coating for Hydrogen Serviced Pipelines.


Coating
Advantage Disadvantage Materials Reference
Type
Various Technique to apply
Limited application
Metal High resistance to erosion Ni-Cobalt, MoS2 [44,48]
to pipelines in use
and corrosion
Brittle
Ceramic Good H2 Resistance Subjected to Thermal SiC, Al2O3, Er2O3 [44,46]
Expansion
HDPE, PE-X,
Swelling Epoxy
Relatively Inexpensive
Blistering Teknopox 3297-
Polymer Good Corrosion Resistance [49]
Hydrogen
Light 00, Teknopox
Permeation
3296-06
Low Permeability Limits on Massive
Two
High Thermal Stability Production
Dimensional MLG, Mxene [47]
Chemically Inert Non-commercialized
Material
High Mechanical Strength method

Metallic coatings [50] are typically used on external pipelines as a sacrificial method to
protect the metal surface through the galvanic principle. These coatings can be applied using
techniques such as hot-dip galvanizing, sherardizing, electroplating, or metal spraying [51]. They
have high resistance to erosive and corrosive environments, as well as atomic hydrogen for certain
metal alloys like nickel alloys and MoS2 [44,48]. However, their use is limited to external surfaces
like bolts and external pipe surfaces, and it is difficult to apply them to the internal wall of pipelines.
Therefore, metallic coatings are not considered suitable for use as an internal barrier coating.
Ceramic coatings are high-temperature coatings based on carbides, silicides, and nitrides
[51]. These coatings are used where chemical resistance and prevention of hydrogen diffusion are
needed [46,51]. While ceramic coatings like SiC, Al2O3, and Er2O3 have high resistance to
hydrogen permeation, they are mechanically brittle and prone to thermal expansion, which can
lead to coating failure under harsh operating conditions. In addition, it is not easy to apply ceramic
coatings inside existing gas pipelines, making them unsuitable for use in existing natural gas
pipelines.
Polymeric coatings are often used for distribution pipeline materials under low pressurized

27
operating conditions due to their relatively low cost, high resistance to corrosion, and low weight
compared to metals. These same advantages also apply to polymeric coatings for pipelines.
According to [52], HDPE and PE-X are suggested as excellent materials for rehabilitating
hydrogen serviced pipelines. Europipe, a German pipeline company, recently announced that the
epoxy-based coatings, Teknopox 3297-00 and Teknopox 3296-06, produced by Teknos, were
suitable for use as internal coatings in pure hydrogen pipelines up to 10 MPa (1,450.38 psi) [53].
While polymeric coatings have multiple advantages, they can be prone to swelling, blistering, and
hydrogen permeation due to poor surface conditions and poor cathodic management [54].
Currently, polymeric coatings are used for corrosion protection in some oil and gas
pipelines worldwide in areas with high corrosion risk. Based on their operational histories [55],
they have had no specific issues in high-pressure pipelines (up to 34.5 MPa or 5,000 psi) for over
20 years. Additionally, it has been reported that using polymeric in-line inspection tools do not
cause abrasion to the internal coating of polymeric coated pipelines [55], which could be a useful
reference for the application of polymeric coatings to hydrogen pipelines.
Two-dimensional materials such as multi-layered graphene (MLG) and Mxene are recent
developments and interesting materials for developing hydrogen barrier coatings due to their
excellent impermeability, chemical inertness, and high mechanical strength. While the large-scale
production of these materials is challenging and the methods for field application have not yet been
developed, they may be considered as promising materials for use as hydrogen barrier coatings in
the future.
It is evident that there are both technical and production challenges to be addressed in the
development of internal coatings for hydrogen transportation. As the demand for hydrogen usage
and transportation grows globally, it is likely that more advanced solutions will be developed, and
existing ones will be further refined. Among the four types of materials currently available,
polymer-based materials appear to be a practical and effective choice for use as internal coatings
in hydrogen pipelines.

4. Hydrogen Leaks
Another consequence with using hydrogen in pipelines (either pure or blended) is the fact
that hydrogen is more prone to leakage compared to methane. To make matters worse, gaseous
hydrogen is colorless and odorless [56], making it difficult to detect with conventional approaches.

28
This also makes it dangerous for operators since humans cannot detect it without special equipment.
Additionally, increasing the operating pressure to improve energy delivery can increase the risk of
leakage (similar to how it interacts with Hydrogen embrittlement, in Section 3).
Leak detection should be in place no matter which type of gas is being transported by
pipelines. This section explores the consequences of hydrogen leaks and potential solutions, such
as the suitability of hardware-based (sensor) methods, Computational Pipeline Monitoring (CPM)
methods and odorants for detecting leaks.

4.1 Hydrogen Leak Behavior


Leakage of hydrogen in an open area is not typically a concern due to its highly diffusive
characteristics. When burning, hydrogen flames release only 10% of the heat produced by
hydrocarbon flames, resulting in minimal damage to the surrounding area [57]. Additionally, the
extremely low density of hydrogen causes it to become buoyant when released into air. Hydrogen
is 14 times lighter than air and 57 times lighter than gasoline vapor [56], making it relatively safe
to work with In open-air spaces as it will usually dilute to non-flammable concentrations quickly
[58].
The use of hydrogen in confined spaces can pose a higher risk due to its wide flammability
range and significantly lower minimum spark energy when compared to natural gas (see Table 6).
This has the potential to cause damage associated with detonation and flammability [59]. If proper
measures and design decisions are not implemented, hydrogen can accumulate quickly in confined
spaces and lead to various hazards, such as initiating a fire, explosion, or even becoming an
asphyxiant at high concentrations.
To address these risks, it is important to ensure proper ventilation when working with
hydrogen in enclosed spaces. A report from the HySafe project recommends ventilation as one of
the most effective safety measures for avoiding dangerous explosive conditions in hydrogen-
treated facilities and provides guidance on the optimal design of a ventilation system in these
facilities [60].
Another potential challenge with hydrogen is the potential stratification of a gas mixture
containing hydrogen due to the significant difference in molecular weights between hydrogen and
other gases. This could potentially lead to increased safety risks, such as in the case of a blended
gas leak event that ends up being primarily composed of hydrogen.

29
The nature of gas stratification in a mixture can vary depending on the flow regime.
Multiple studies note that blended gases tend to mix well in turbulent flow [22,61]. A particular
study noted that the mixture becomes homogenous within a short distance from the mixing point
[11]. In contrast, laminar flow initially stratifies the mixture, but it becomes homogenous after a
certain length. This critical distance can be roughly estimated by multiplying the nominal diameter
by 4,000 [61]. Therefore, it is important to pay particular attention to operating pipelines with
lower Reynolds numbers to minimize the risk of gas stratification.
Mejia et al. [62] conducted leak tests with blended hydrogen to assess the difference in leak
rates of each gas component in the mixture. The tests were conducted with fully mixed 95% natural
gas and 5% hydrogen at pressures up to 417 kPa (60.5 psi). The results showed that the difference
in leak rates for each gas was minimal (around 1%). While these results are encouraging, it is
important to note that they were conducted at relatively low pressure and started from a fully mixed
state.
Another study examined the relative size of gas leaks under different flow regimes (laminar,
turbulent, and sonic), with leak rates estimated using the dynamic viscosities, molar masses, and
gas densities of the gases [63]. The results, shown in Table 12, indicate that the estimated leak rate
was significantly higher for hydrogen in all cases. For example, in the case of sonic flow (typical
of leaks in high-pressure lines), hydrogen would leak volumetrically 2.8 times more than methane.
However, when considering energy density, the energy contained in such a hydrogen leak would
be 0.88 times a similar methane leak.

Table 12. Leak Rate by Different Types of Flow Regimes. [63]

Properties Hydrogen Methane Unit

Subsonic Flow (Laminar) 1.3 1


Subsonic Flow (Turbulent) 2.8 1 Relative leak rates (Volumetric)
Sonic Flow 2.8 1

4.2 Sensor Based Hydrogen Leak Detection


One concern with using traditional hardware-based leak detection methods for blended
hydrogen gas is their potential compatibility issues due to the high mobility, small size, and high

30
diffusivity of the gas with other materials. While further testing is needed to determine the
effectiveness of these methods with blended hydrogen gas, some studies suggest that it could be
solved with minor sensor recalibration [10]. It is important to evaluate the reliability and
compatibility of hydrogen leak detection systems for each specific operating scenario.
Alternatively, specialized leak detection facilities designed specifically for hydrogen may
be a viable option for detecting leaks of blended hydrogen gas. There are four main types of sensors
used for hydrogen detection: semiconductor metal oxide, electrochemistry, catalytic bead, and
thermal conductivity [64–66]. These are summarized in Table 13.

Table 13. Comparison of Commercially Available Hydrogen Detecting Sensors [64–66].


Measuring Response
Sensor Type Accuracy Cost Features
Range Time (t90)
- Low Cost
Semiconductor 0 – 1000 $100 -
± 10-30 % <20 /s - Dependence on Humidity
Metal Oxide ppm $500
and Temperature
- Good Selectivity to
0 – 20000 $300 -
Electrochemical < ± 4% <90 /s Hydrogen
ppm $1200
- Narrow Temperature Range
$500 - - Wide Temperature Range
Catalytic Bead < ± 5% 0 – 100 % H2 <30 /s
$4,000 - No Hydrogen Selectivity
Thermal - High Accuracy
± 0.2% 0 - 100% H2 <10 /s <$25,000
Conductivity - Cross Sensitive to He

Semiconducting metal oxide (MOx) sensors are widely used in industry due to their
economic nature, but they have the disadvantage of low accuracy and sensitivity to humidity and
temperature [64,67]. Electrochemical sensors offer good selectivity for hydrogen gas and relatively
high accuracy, but they are limited in their temperature range and have slow detection times.
Catalytic sensors have a wide range of hydrogen detection in volume and operating temperature,
but they are expensive and not highly selective for hydrogen gas.
Thermal conductivity sensors have the broadest measurement range and the highest
detection accuracy among the four sensor types in Table 13. However, they are prone to cross-
sensitivity readings with helium (He). Given the importance of economic and agile considerations
for industrial applications, semiconductor-based hydrogen leak detection methods may be a
realistic option.
Due to the high diffusive nature of hydrogen, the location of sensors is critically important

31
to accurately detect leaks. Such sensors may be more appropriate for confined spaces. Moreover,
further study is needed to investigate how these sensors would behave for blended hydrogen and
natural gases.

4.3 Computational Pipeline Monitoring (CPM) Methods


CPM methods involve using a computer to monitor changes in measured pipeline data for
leak identification in pipeline networks. This method often involves the use of a supervisory
control and data acquisition (SCADA) system, which collects sensor data from pipelines, transmits
it to computers, and uses data processing to identify leaks. The most reliable and widely used CPM
system is the real-time transient model (RTTM), which simulates pressure and flow rate in
pipelines using actual sensor data [66]. The RTTM compares the simulated data to the actual sensor
data to identify leaks. While it has been successfully implemented and is widely used for natural
gas pipelines, it has not yet been proven for use in hydrogen gas pipelines.
A simple real-time transient model (RTTM) is developed to verify whether it can be used
for blended hydrogen gas pipelines. The model is based on the simulation parameters introduced
in Section 2, and it includes conservation equations for mass, momentum, and energy. These
partial differential equations (PDEs) are solved using the method of characteristics (MOC) [68].
The typical conservation equations in PDE form are shown in Eq. (13) and Eq. (14).

𝜕𝐻 𝑎𝑠2 𝜕𝑄
+ =0 (13)
𝜕𝑡 𝑔𝐴 𝜕𝑥
1 𝜕𝑄 𝜕𝐻 𝑓𝑄|𝑄|
+ + =0 (14)
𝑔𝐴 𝜕𝑡 𝜕𝑥 2𝑔 ⋅ 𝐷 ⋅ 𝐴2
where H is the hydraulic head (m), Q is the volumetric flow rate (m3/s), g is the gravitational
acceleration (m/s2), a is the speed of sound (m/s), f is the Darcy friction factor (dimensionless), D
is the internal diameter of the pipe (m), A is the cross-sectional area of the internal pipe (m2), and
t is time (s).

Since these equations are time- and distance-dependent, the flow velocity can be estimated
using time and distance. This allows for the estimation of hydraulic information in pipes at specific
points in time and space, given the size of the time step (𝜕𝑡) and distance step (𝜕𝑥). The calculation

32
procedure is well-established in the literature [68,69] and involves applying the general gas flow
equation (Eq. (1)) to calculate pressure and flow rate. The temperature-dependent parameters
needed for this calculation, including density, viscosity, specific gravity, and compressibility for
hydrogen and methane gas mixtures, are obtained from GERG-2008, a wide-range equation of
state [70,71].
Leakage is simulated by characterizing the pressure difference between the internal
pressure of the pipeline and the atmospheric pressure. The discharge coefficient, which accounts
for energy loss due to friction and compressibility effects, is also taken into consideration [72].
The energy conservation law, as described by Bernoulli's equation (Eq. (15)), can be simplified to
Eq. (16) by assuming that the pipeline has a constant elevation profile and that the leak flow rate
from the leak hole is significantly smaller than the flow rate inside the pipe:

𝑃𝐴 𝑉𝐴2 𝑃𝐵 𝑉𝐵2
ℎ𝐴 + + = ℎ𝐵 + + (15)
𝛾𝑤 2𝑔 𝛾𝑤 2𝑔

2∆𝑃
𝑄 = 𝐶𝑑 ⋅ 𝐴√ (16)
𝑆𝑔 ⋅ 𝜌

where h is the elevation (m), P is the pressure (Pa), V is the flow velocity (m/s), 𝛾𝑤 is the specific
weight (N/m3), Cd is the discharge coefficient (dimensionless), 𝑆𝑔 is the specific gravity
(dimensionless), and 𝜌 is the density (kg/m3).
To consider turbulent flow, a discharge coefficient of 0.67 is chosen as this is typical for
high Reynolds number flow [72]. The simulation is defined by two boundary conditions: a pressure
boundary at the upstream side and a flow velocity at the downstream side. The leak is simulated
as a relative size of 10% of the pipe's internal diameter. The simulation is carried for a total of
2000 seconds, with a leak event happening at the center (1000 seconds). Figure 11 shows the
simulation results for a range of hydrogen concentrations (from 10% to 90% hydrogen mixed with
methane) in the presence of a leak.
The figure shows the impact of a potential leak on the simulated pipeline, with pressure
shown in (a) and flow velocity in (b). It's worth noting that the pressure starts at the same level for
all gas mixtures due to the pressure boundary condition. It's also worth noting that, as flow velocity
is density-dependent, the different gas mixtures behave differently both up and downstream.
The pressure drop in each gas mixture shows similar, but distinct trends due to the

33
differences in compressibility and friction factor of the hydrogen components in the gas mixture
(Eq. (1)). Flow velocity has similar trends in all simulation cases, showing an increase in velocity
after the leakage. This increase in velocity is due to the lower pressure caused by the leakage in
the system [34].

(a) Upstream Pressure with Leakage (b) Downstream Flow Velocity with Leakage
Figure 11. Pressure and Flow Velocity Variation with Leakage by Hydrogen Concentration.

In summary, the model developed in this study shows that while the gas mixtures have
small differences in their leak behavior, they also have very similar trends. While the RTTM model
is relatively simple and may not be directly applicable to complex real-life pipeline networks, it
suggests that with proper calibrations, RTTM models can be utilized for blended gas mixtures.

4.4 Odorization
Another potential approach to hydrogen leak detection is gas odorization, which involves
adding chemical odorants to the gas that can be detected by the human olfactory sense. This
method is already widely used for natural gas, but the odorants used for natural gas are not suitable
for hydrogen due to their significantly different densities and dispersion rates. While there are
currently no known commercially usable odorants for hydrogen, several studies have evaluated
odorants as a potential method for detecting hydrogen.
A research group in the Netherlands, consisting of Gasunie Transport Services (GTS) and
Netbeheer, evaluated three odorants (THT, Spotleak 1001, Gasodor S-Free) at hydrogen

34
concentrations ranging from 0% to 100%. They concluded that these odorants could potentially be
used for a mixture of natural gas and hydrogen in distribution lines [25].
The Hy4Heat project in the United Kingdom also evaluated the hydrogen odorization
performance of five chemical compounds across six different sectors: Health and Environment,
Olfactory, Pipeline, Flame Boiler, Fuel Cell, and Economic [25]. According to the evaluation
results, the five odorants can be used for pure hydrogen except in fuel cell applications due to
sulfur poisoning. The potential odorants for hydrogen gas are summarized in Table 14.
The odorants selected in this table were observed to be non-corrosive and metallurgically
intact. Their functionality was also proven through an olfactory test that evaluated their
detectability at a minimum of one-fifth of the lower flammability limit of the gas composition and
compatibility with hydrogen. In addition, the dilution effect by the addition of odorants is not
foreseen for the listed odorants. All the odorants can potentially be applied to hydrogen serviced
pipelines for general-purpose (but not necessarily fuel cell applications). Spotleak 1001 is
particularly suitable as a hydrogen odorant as it is broadly used in European gas networks and is
low cost with no harmful effects on pipelines or other appliances [25].

Table 14. The List of Odorants Studied for Use with Hydrogen Gas. [25]
Olfactory Material Voltage Fuel Cell
Name (Compound) Sulfur Health Corrosive*7
Test Degradation Loss(mv)*3 Application
Purification
Spotleak 1001*1 × × Pass × × 460 ±2
Required
Standby Odorant 2*2 × × Pass × × 40±2 O
*3 Purification
THT × × Pass × × 225±2
Required
Gasodor-S-Free*4 × × Pass × × 10±2 O
Norbornene*5 × × ×*6 × × 5±2 O
1*
Spotleak 1001: 2-Methyl-propanethiol 78%, Dimethyl Sulfide 22%
2*
Standby Odorant2: Odorant NB 34%, Hexane 64%
3*
Odorant THT: Tetrahydrothiophene – 100%
4*
Gasodor-S-Free: Methyl Acrylate 37.4%, Ethyl Acrylate 60.1%, 2-Ethyl-3-Methylpyrazine 2.5%
5*
Norbornene: 5-Ethylidene-2-Norbornene
6*
Not distinguishable with sulphur and oil
7*
After Complete Combustion

35
5. Hydrogen Influence on Greenhouse Gas Emissions
One of the main benefits of using hydrogen as an energy carrier is its ability to reduce
greenhouse gas (GHG) emissions from fossil fuel combustion. While the combustion of hydrogen
can produce nitrogen oxides (NOx) at high temperatures [9], it is otherwise virtually emission-free
[73]. This is particularly important in light of increasing concern over GHG emissions being
produced globally. As such, the use of hydrogen as an energy carrier can help to offset these
emissions and promote a more sustainable energy system.
CO2 is one of the main GHGs responsible for global warming, along with methane (CH4),
nitrous oxide (N2O), and F-gases. CO2 is the most prevalent GHG, accounting for 76% of global
GHG emissions [67]. To compare the global warming potential (GWP), they are often expressed
in terms of equivalent CO2 [74,75]. Table 15 shows the primary GHGs expressed in terms of
equivalent CO2 for carbon dioxide, methane, N2O, and HFC emissions.

Table 15. GWP for Each Gas in 1970 and 2021. [75]

1970 2010
GHG Emissions
GtCO2 – eq/yr % GtCO2 – eq/yr %

Total 19.44 72 37.24 76


Forestry and
4.59 17 31.85 11
CO2 Other Land Use
Fossil Fuel and
Industrial 14.85 55 5.39 65
Processes
CH4 5.13 19 7.84 16
N2O 2.133 7.9 3.038 6.2
F-Gases 0.1188 0.44 0.98 2
Total 27 ≈ 100 49 ≈ 100

In the following we examine the effect of hydrogen blending on greenhouse gas emissions.
We first provide an overview of various methods of hydrogen production and conversion into
power. Next, we use a realistic simulation to determine the potential reduction in GHG emissions
that could be achieved by converting a natural gas pipeline (described in Section 2.2) into a blended
hydrogen pipeline using currently available technology (blue hydrogen). It should be noted that

36
while the ideal scenario would be for most hydrogen to be produced through renewable sources
(green hydrogen), it is at present not feasible to rely solely on such methods.

5.1 Hydrogen Production


Hydrogen production can be classified into four categories based on the carbon emissions
and production methods: Blue, Gray, Green, and Purple hydrogen. Blue hydrogen is produced
through carbon capture and storage (CCS) methods, while Gray hydrogen is produced from fossil
fuels through steam methane reforming (SMR) or autothermal reforming (ATR) methods, which
generate carbon dioxide (CO2). Green hydrogen is produced through the environmentally friendly
process of water electrolysis using renewable energy, resulting in minimal carbon emissions during
production. Purple hydrogen is produced through electrolysis powered by nuclear energy. Another
potential technology for low-carbon hydrogen production is Proton hydrogen, which involves the
use of a membrane-based separation applied to oil reservoirs to selectively filter hydrogen gases
out of fossil fuel-based gas mixtures or vapors. The advantages and disadvantages of each type of
hydrogen production are summarized in Table 16.

Table 16. Hydrogen Generation Comparison. [16,76–79]


Type Advantages Disadvantages
Less Carbon Emission (≈ 90%)
Blue H2 Low Production Efficiency
Mature Technology
Low Production Cost
High Carbon Emissions
Grey H2 (2~3 dollars/kg)
(No Carbon Capture Process)
Mature Technology
High Cost (>$5/kg H2)
Zero Carbon Emission Low Production Efficiency
Green H2
Simple Method Discontinuous Electrolysis process
High Initial Cost for Infrastructure
Low Carbon Emission
Use for Surplus of Energy for
hydrogen production (electricity Safety Concerns with Nuclear
Purple H2
and thermal energy from reactor) Relatively Unmatured Technology
Continuous hydrogen production
Large Production
Low Carbon Emission
Massive Production Limited to Geological location
Proton H2
Economical Production Immature Technology
(Below 30 cents/kg)

37
Low-Cost Installation

Low carbon-intensity hydrogens, such as Green, Purple, or Proton hydrogen, are important
for achieving long-term carbon-neutralization in energy consumption. However, producing these
types of hydrogen economically can be challenging due to the high cost of alkaline electrolyzers
and low production efficiency [4]. While the technology is expected to become more cost-effective
as it matures, estimates suggest that the current cost of hydrogen could potentially be reduced from
the range of $2.5 to $6.64 per kilogram to the range of $0.7 to $2.6 per kilogram by 2050 [5,80,81].
Additionally, the production of Green or Purple hydrogen requires significantly more water than
Blue, Gray, or Proton hydrogen.

5.2 Hydrogen to Power


Hydrogen can be converted into power through two primary means: combustion in internal
combustion engines, heating systems, and appliances, or through an electrochemical reaction in
fuel cells to generate electricity. While hydrogen can be burned with oxygen to produce heat
without emitting carbon when used as a combustion fuel, burning hydrogen with air at high
temperatures above 1,500°C can produce the undesirable byproduct of nitrous oxide (NOx) [9].
Despite this, hydrogen remains a promising alternative to fossil fuels for reducing carbon
emissions.
Hydrogen has the potential to serve as a carbon-free alternative to natural gas in various
industrial applications, such as heating and end-use appliances. However, it is important to note
that conventional gas appliances will most likely need to be retrofitted in order to be compatible
with pure hydrogen. An alternative option is to use blended hydrogen, which can be used in
existing gas-based systems without modification or with only minor modifications [82]. Table 17
summarizes the results of multiple studies on the hydrogen tolerance of various appliances with
combustion processes based on the concentration of hydrogen used. This shows that most
combustion appliances are compatible with blended hydrogen at concentrations up to 25 % in
volume.

Table 17. Hydrogen Tolerance for Combustion Appliances.

Type H2 Tolerance Method of Theory Interchangeability References

38
Up to 15% Wobbe Index, Flame Speed, NOx
Gas Turbine [83]
(Minor changes) Emission
Up to 10% Wobbe Index, Flame Speed, NOx
Gas Engines [83]
(Minor changes) Emission
Ignition Performance, Flame
Water Heater 10% [84]
Characteristics, Emission
Cooking Performance, Ignition Time,
Cooktop Burner 20 vol% Flame Characteristics, Combustion [32]
Noise, Burner Temperature, Emission
Oven Burner 25 vol% Wobbe Index [85]
End-use Appliance
17 vol% Wobbe Index [36]
(Burner, Boiler,)

Hydrogen has the potential to serve as a carbon-free alternative to natural gas for electrical
power generation using a cleaner and more efficient process than combustion. Fuel cells typically
generate energy at relatively low temperatures, which prevents the formation of the previously
mentioned NOx. Fuel cells can achieve high efficiencies of between 60% and 70% [9] and are not
bound by the Carnot efficiency [9] of traditional heat engines. Fuel cells are generally classified
based on their temperature range and electrolyte, as summarized in Table 18.

Table 18. Comparison of Different Types of Fuel Cells. [9]


Temperature Life
Fuel Cell Type Electrolyte Efficiency Cost
Range Expectancy
Alkaline Fuel Potassium 5,000 – 200 - 700
60 – 90°C 50 – 60%
Cell (AFC) Hydroxide 8,000 Hr USD/kWe
Proton Exchange
Polymer
Membrane Fuel 50 - 180°C 30 – 60% 5,000 Hr 500 USD/kWe
Membrane
Cell (PEMFC)
Phosphoric Acid Phosphoric 30,000 - 4,000 -5,000
160 - 220°C 30 – 40%
Fuel Cell (PAFC) Acid 60,000 Hr USD/kWe
Motel Carbonate 20,000 - 4,000 -6,000
600 - 700°C Carbonate Melt 55 – 60%
Fuel Cell (MCFC) 40,000 Hr USD/kWe
Oxide Ceramic Solid Ceramic Up to 3,000 -4,000
700 - 1,000°C 50 – 70%
Fuel Cell (SOFC) Oxide 90,000 Hr USD/kWe

The challenge with the widespread deployment of fuel cells in a hydrogen economy that
relies on blended hydrogen, however, is that most electrolyzers require different degrees of
hydrogen purity. For example, the International Organization for Standardization (ISO) and SAE
international issued technical standards (ISO14687:2019, SAE J2719-202003) [86,87] for the

39
hydrogen quality required for proton exchange membrane fuel cells [88] and the hydrogen purity
necessary for fuel cells. The requirements are described in Table 19.

Table 19. Hydrogen Purity Requirements for Fuel Cells by ISO and SAE Standards. [88]
Non
Contents H2 He N2 Ar CH4 H2O Others Total Sulfide
H2 Gas
Requirement 99.97% 300ppm 300ppm 300ppm 300ppm 100ppm 5ppm 2.75ppm 0.004ppm

Separation of hydrogen from natural gas is not trivial, but there are already three
commercially used and well-stablished methods: Pressure Swing Absorption (PSA), Membrane
Systems and Cryogenic Distillation. These are summarized in Table 20.

Table 20. Types of Separation Technologies. [10,89]


Cryogenic
Category PSA Membrane
Distillation
Input Composition
30-75 75-90 30-90
(H2 mol%)
Output Purity
90-98 95-99 90-98
(H2 mol%)
Processing Volume
>10,000 1000-10,000 <30,000
(Nm3/h)
Reliability (%) Poor 95 100
Turndown (%) 10 30 30-50
Large Separation
- -
Facilities
Size
High Pressure
High Pressure Drop -
Required
Applicability Distribution Line Transmission Line -

PSA is a well-established technology that is used in refineries to produce large quantities


of high-purity hydrogen (90-98%). This process involves applying a pressure drop to separate
hydrogen from impurities, and thus the most suitable location for applying hydrogen separation
using PSA technology is at pressure reduction stations, where a high pressure drop occurs. It is
capable of producing hydrogen at rates ranging from 50 to 200,000 Nm3/h [10] from input gases
containing 30 to 75 mol% hydrogen. One advantage of PSA is its low operating cost and long
service life, but it can also have relatively higher hydrogen losses and a lower production rate due

40
to the high purity requirements for hydrogen [88].
Membrane technology is an effective method for obtaining high purity hydrogen (95-99%)
and is capable of large-scale processing. It works by utilizing the pressure gradient across a
membrane. One of the main advantages of this technology is its reliability in the case of unexpected
shutdowns, as it has no mechanical parts [88]. However, its main disadvantage is that it requires
an input with a fairly high concentration of hydrogen (75-90%), which can pose challenges for
pipelines that transport lower concentrations.
Cryogenic distillation is another process used in large-scale hydrogen recovery from fluid
mixtures, which utilizes very low temperatures to separate the gas mixture based on the difference
in volatility of each gas. However, this method has its own challenges, such as the need to filter
impurities such as carbon dioxide and water content before separation [88], and the high energy
consumption required to compress and cool down the gas mixture. These challenges make it
difficult to produce sustainable and economical hydrogen for general purposes.

5.3 Influence on GHG Emissions


In order to assess the influence of hydrogen concentrations and energy transmission on GHG
emissions, the following sections outline the mathematical equations governing emissions related
to combustion and production. These equations are then combined and applied to the simulation
model developed in Section 2. The resulting visualization provides a comprehensive depiction of
the impact on GHG emissions.

5.3.1 From Combustion

To calculate the GHG Emissions from the combustion of hydrocarbons the complete
combustion process is considered as balanced with regards to oxygen, and thus, all carbon-
containing fuel is consumed and converted into carbon dioxide and water. The governing equation
is shown in Eq. (17). The variables x, y, and z are the stoichiometric coefficients for carbon,
hydrogen, and oxygen, respectively.

𝑦 𝑧 𝑦
𝐶𝑥 𝐻𝑦 𝑂𝑧 + (𝑥 + − ) 𝑂2 → (𝑥)𝐶𝑂2 + ( ) 𝐻2 𝑂 (17)
4 2 2

41
The amount of CO2 released by combustion can be calculated by multiplying the fuel
consumed with the mass percentage of the fuel's carbon content. For fuels consisting of a single
component, the carbon content of this component can be obtained with the following mathematical
relation:

12 × 𝑥
𝐶𝑠 = × 100% (18)
𝑀𝑤

where Cs is the carbon content of the component (mass %), 12 is the molecular weight of carbon,
x is the stoichiometric coefficient for carbon, and Mw is the molecular weight of the component.

For mixed gases, the carbon content of the mixture (𝐶𝑚 ) can be obtained by multiplying
the mass percent of each gas in the total gas mixture after calculating the carbon mass percentage
of each fuel, as shown in the following equation.

𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝐶𝑜𝑚𝑝𝑜𝑛𝑒𝑛𝑡𝑠
1
𝐶𝑚 = × ∑ (𝑊𝑡%𝑖 × 𝐶𝑠 𝑖 ) (19)
100
𝑖=1

where Cm is the carbon content of mixture (Mass %), 𝑊𝑡%𝑖 is the weight fraction of component
©, and 𝐶𝑠 𝑖 is the carbon content of component © on a mass percent basis.

Finally, the equivalent CO2 emission (𝑬𝑪𝑶𝟐 ) can be estimated by multiplying the carbon
mass content with the total amount of fuel and the molar mass of the gas mixture [74]. This is
shown in Eq. (20):

𝟏 𝟒𝟒
𝑬𝑪𝑶𝟐 = 𝑭𝑪 × × 𝑴𝒘 × 𝑪𝒎 × (20)
𝑴𝒐𝒍𝒂𝒓 𝒗𝒐𝒍𝒖𝒎𝒆 𝒄𝒐𝒏𝒗𝒆𝒓𝒔𝒊𝒐𝒏 𝟏𝟐
where 𝐸𝐶𝑂2 is the emission of CO2 in mass (kg), FC is the volume of fuel consumed (m3), Molar
volume conversion is the conversion value from molar volume to mass (23.685 m3/kg mol), 𝑀𝑤 is
44
the molecular weight of the mixture, and 12 is the stoichiometric conversion from carbon to CO2.

42
5.3.2 From Hydrogen Production

To develop a realistic model, the use of hydrogen in the simulation is assumed to be


produced through the SMR process, that is, as Blue hydrogen. This process involves catalytically
reacting natural gas with steam at high temperatures to produce hydrogen. However, it should be
noted that the production of Blue hydrogen generates CO2 emissions due to the heat required for
the reforming process being obtained through the combustion of fossil fuels [90]. The following
chemical reactions represent the general process of SMR:

𝐶𝐻4 + 𝐻2 𝑂 → 𝐶𝑂 + 3𝐻2 (21)

𝐶𝑂 + 𝐻2 𝑂 → 𝐶𝑂2 + 𝐻2 (22)

The SMR reaction produces a molar ratio of 0.25 for CO2 and H2, meaning that for every
1 kg of CH4 reacted, 1 kg of CO2 and 4 kg of H2 are produced under ideal conditions. However, in
the actual SMR process there are tail gases from the pressure swing absorption (PSA) stage, which
consist of unreacted methane and some hydrogen gas. These tail gases can be utilized to fuel the
reformer and boiler units. When accounting for this gas recovery process, the molar ratio between
CO2 and H2 improves from 0.25 to 0.395 [91]. Finally, the simulation considers the use of a carbon
capture and storage (CCS) stage, where emissions from hydrogen production can be mostly
compensated for. According to the CE delft report [92], the CCS stage is estimated to have a
performance of 90%.

5.3.3 Combined Impact of Combustion and Production

The model developed in Section 2 is used in conjunction with the previous equations. It is
assumed that the entire amount of the gas mixture being transported is consumed for heating or
electricity generation through the use of turbines and a complete combustion process. Figure 12
shows the expected GHG emissions as a function of hydrogen concentration and energy
transmission. In (a) the influence of the combustion process is shown independently, while (b)
illustrates the influence of the SMR process (hydrogen production) independently. Finally, (c)
shows their combined influences.

43
(a) GHG Emissions in Combustion (b) GHG Emissions in Production
(Logarithmical Z-scale) (Linear Z-scale)

(c) © Combined GHG Emissions


(Logarithmical Z-scale)
Figure 12. GHG Emissions as Factors of Hydrogen Concentration and Energy Transmission.

In (a), it is shown that a higher concentration of hydrogen leads to a reduction in CO2


emissions during the combustion process. In fact, pure hydrogen usage results in no CO2 emissions
at all. In contrast, the highest CO2 emissions are found when using pure methane (natural gas) at
the initial flow rate.
In (b), it is demonstrated that pure hydrogen with 100% transmitted energy tends to emit
the highest amount of CO2. This is due to the higher flow rates required to match the transmitted
energy, as well as the increased production of hydrogen to meet this flow. It is also shown that

44
CCS greatly mitigates these emissions.
In (c) the full picture is shown. When combining the influence of both the combustion and
SMR processes, it becomes clear that the increase in GHG emissions by the SMR process is greatly
offset by the reduction in GHG emissions through the combustion process, making the emissions
for hydrogen production almost negligible.
The baseline for CO2 emissions is represented by the black shaded square, which
corresponds to the initial flow rate (14,158,423 m3/day). Additionally, red arrows indicate the
reduction on GHG emissions for different scenarios. For example, it can be seen that in extreme
cases, the use of pure hydrogen at either 100% or 30% transmitted energy can lead to a reduction
of 99.4% or 99.7% in CO2 emissions, respectively.
A less extreme example can be considered by adding 30% hydrogen to the gas mixture at 51%
energy transmission, one of the potential solutions discussed in Table 7. In this case, the GHG
emissions can have a reduction of 36.3%, or a decrease of 1,127,233 ton/day. While it would be
ideal to see larger changes and higher concentrations of hydrogen, it may not be feasible at this
time. It is worth noting that even a 30% blend has significant potential for reducing emissions and
is achievable with current technology.

6. Conclusions
One of the economically viable ways to transport and distribute hydrogen is by blending it
with natural gas using existing the pipeline networks. However, it has been shown that blending
hydrogen can reduce the energy transmission of pipelines compared to natural gas. One potential
solution is to increase the flow rate or pressure of delivery to compensate for these energy
transmission losses, which may require changes to the pipeline or increase the number of
compressor stations. In the studied pipelines, it was found that hydrogen concentrations of up to
30% with a 49% reduction in energy transmission could be immediately applied without any
changes. If a compressor upgrade is budgeted, a blend of up to 60% hydrogen with a 35% reduction
in energy transmission could be achieved.
In addition to energy considerations, it is important to consider the potential risks of
hydrogen embrittlement and leak events when increasing pressure. However, many materials have
been tested for their susceptibility to hydrogen embrittlement, and best practices for dealing with
this risk are available. For materials that may not be compatible with hydrogen, internal coating

45
can be applied to improve resistance to hydrogen embrittlement. Leak detection is also a concern,
but hardware sensing options and CPM (Computational Pipeline Monitoring) methods are likely
to be compatible with leaks in hydrogen systems requiring only minor calibration. To demonstrate
this, a RTTM (Real-Time Transient Model) that utilizes the same principle used in natural gas
systems to detect pressure and flow velocity changes was developed for hydrogen blends.
One of the main goals of hydrogen blending and the hydrogen economy is to reduce
greenhouse gas (GHG) emissions. In order to achieve this, a model was developed that examined
the impact of adding Blue hydrogen (produced via steam methane reforming with carbon capture
and storage) to existing pipelines without requiring any changes to the pipelines themselves. The
model showed that blending 30% hydrogen with the current pipeline system would result in a 49%
reduction in energy transmission, and a 36.3% reduction in GHG emissions if we assume that the
blended hydrogen will be used for heating purposes.
These findings indicate that even though Blue hydrogen production does produce CO2
emissions, it is still possible to reduce GHG emissions overall through its usage. Blended hydrogen,
which can be implemented using current technologies, may be a useful compromise until we
migrate to pure hydrogen in the future. This study aims to provide practical insights into the
potential benefits and challenges of using blended hydrogen in existing pipelines and contribute to
the current understanding of this topic.
In future studies, it is expected that progress will be made in addressing the challenges
mentioned, such as hydrogen embrittlement and improving hydrogen specific leak detection. This
may include the development of new strategies and technologies for mitigating the effects of
hydrogen embrittlement and detecting leaks more effectively.
It would also be valuable to examine the economic feasibility of implementing blended
hydrogen in existing pipeline networks, including the costs of necessary upgrades or modifications.
This could involve modeling the potential savings from reduced GHG emissions and improved
energy transmission, as well as the potential for revenue from hydrogen when compared to natural
gas.
Another interesting area of study could be the examination of the potential benefits and
challenges of using ammonia as a hydrogen carrier and comparison to the direct use of hydrogen.
Ammonia has a high hydrogen content, making it a potential option for the transportation and
distribution of hydrogen. However, its production, storage, and conversion to and from hydrogen

46
must also be considered, as well as the requirements for safe handling. These factors should be
weighed against the potential benefits of using ammonia, such as its ability to be easily liquefied,
in order to determine its feasibility as a hydrogen carrier.

Acknowledgements
The authors would like to thank the Natural Sciences and Engineering Research Council
of Canada (NSERC), Alberta Innovates, Mr. Kalen Jensen from ATCO and the Institute for
Information & Communication Technology Promotion (IITP) for their support of this research
project.

References
[1] Kamarudin, S. K., Daud, W. R. W., Yaakub, Z., Misron, Z., Anuar, W., & Yusuf, N. N. A. N.
(2009). Synthesis and optimization of future hydrogen energy infrastructure planning in Peninsular
Malaysia. International Journal of Hydrogen Energy, 34(5), 2077–2088.
https://doi.org/10.1016/j.ijhydene.2008.12.086
[2] Mazloomi, K., & Gomes, C. (2012). Hydrogen as an energy carrier: Prospects and challenges.
Renewable and Sustainable Energy Reviews, 16(5), 3024–3033.
https://doi.org/10.1016/j.rser.2012.02.028
[3] Götz, M., Lefebvre, J., Mörs, F., McDaniel Koch, A., Graf, F., Bajohr, S., Reimert, R., & Kolb,
T. (2016). Renewable Power-to-Gas: A technological and economic review. Renewable Energy,
85, 1371–1390. https://doi.org/10.1016/j.renene.2015.07.066
[4] Oil & Gas Authority. (2018). Gas to Wire report: UK SNS and EIS.
https://www.ogauthority.co.uk/news-publications/publications/2018/gas-to-wire-repor t-uk-sns-
and-eis/
[5] Bloomberg New Energy Finance (BNEF). (2021). New Energy Outlook 2021.
https://about.bnef.com/new-energy-outlook/
[6] Canadian Energy Research Institute (CERI). (2021). Canadian Natural Gas Market Supply and
Demand Pathways of Change. CERI, Calgary, AB.
[7] Apostolou, D., & Xydis, G. (2019). A literature review on hydrogen refuelling stations and
infrastructure. Current status and future prospects. Renewable and Sustainable Energy Reviews,
113, 109292. https://doi.org/10.1016/j.rser.2019.109292
[8] Veziroğlu, T. N., & Şahi˙n, S. (2008). 21st Century’s energy: Hydrogen energy system. Energy
Conversion and Management, 49(7), 1820–1831. https://doi.org/10.1016/j.enconman.2007.08.015

47
[9] Adolf, J., Balzer, C. H., Balzer, C., Louis, J., Schabla, U., Fischedick, M., Arnold, K., &
Pastowski, A. (2017). Shell Hydrogen Study. Energy of the future? Sustainable Mobility through
Fuel Cells and H2. Shell Deutschland Oil GmbH.
[10] Melaina, M. W., Antonia, O., & Penev, M. (2013). Blending Hydrogen into Natural Gas
Pipeline Networks: A Review of Key Issues. Renewable Energy, 131.
[11] Isaac, T. (2019). HyDeploy: The UK’s First Hydrogen Blending Deployment Project. Clean
Energy, 3(2), 114–125. https://doi.org/10.1093/ce/zkz006
[12] H21. (2019). H21 Leeds City Gate Report. Westminster, UK. https://h21.green/projects/h21-
leeds-city-gate/
[13] Blokland, H., Sweelssen, J., Isaac, T., & Boersma, A. (2021). Detecting hydrogen
concentrations during admixing hydrogen in natural gas grids. International Journal of Hydrogen
Energy, 46(63), 32318–32330. https://doi.org/10.1016/j.ijhydene.2021.06.221
[14] International Energy Agency. (2021). Global Hydrogen Review 2021. OECD.
https://doi.org/10.1787/39351842-en
[15] Australian Gas Infrastructure Group. (2019). Renewable Hydrogen, Hydrogen Park SA.
World Plumbing Conference. https://www.worldplumbing.org/wp-content/uploa
ds/2020/06/VikramSingh_RenewableHydrogenHydrogenParkSA.pdf
[16] Elmanakhly, F., DaCosta, A., Berry, B., Stasko, R., Fowler, M., Wu, X.-Y., & Hydrogen
Business Council. (2021). Hydrogen economy transition plan: A case study on Ontario. AIMS
Energy, 9(4), 775–811. https://doi.org/10.3934/energy.2021036
[17] Ozturk, M., & Dincer, I. (2021). A comprehensive review on power-to-gas with hydrogen
options for cleaner applications. International Journal of Hydrogen Energy, 46(62), 31511–31522.
https://doi.org/10.1016/j.ijhydene.2021.07.066
[18] Republica Portuguesa. (2020). Portugal National Hydrogen Strategy (EN-H2).
https://kig.pl/wp-content/uploads/2020/07/EN_H2_ENG.pdf
[19] Saeedmanesh, A., Mac Kinnon, M. A., & Brouwer, J. (2018). Hydrogen is essential for
sustainability. Current Opinion in Electrochemistry, 12, 166–181.
https://doi.org/10.1016/j.coelec.2018.11.009
[20] Zhao, Y., McDonell, V., & Samuelsen, S. (2019). Influence of hydrogen addition to pipeline
natural gas on the combustion performance of a cooktop burner. International Journal of Hydrogen
Energy, 44(23), 12239–12253. https://doi.org/10.1016/j.ijhydene.2019.03.100
[21] Krieg, D. (2012). Konzept und Kosten eines Pipelinesystems zur Versorgung des deutschen
Straßenverkehrs mit Wasserstoff. Forschungszentrum Jülich.
[22] Kong, M., Feng, S., Xia, Q., Chen, C., Pan, Z., & Gao, Z. (2021). Investigation of Mixing
Behavior of Hydrogen Blended to Natural Gas in Gas Network. Sustainability, 13(8), Article 8.
https://doi.org/10.3390/su13084255

48
[23] ATCO Gas Australia. (2019). Clean Energy Innovation Hub Lessons. Arena Insights Forum.
https://arena.gov.au/assets/2019/12/atco-clean-energy-innovation-hub-lessions.pdf
[24] GRTgaz, Immeuble Bora, 6, rue Raoul-Nordling, 92277 Bois-Colombes Cedex. (2019).
Technical and economic conditions for injecting hydrogen into natural gas networks—Final report
June 2019. France, INIS-FR--20-0156. https://www.afgaz.fr/wp-content/uploads/Technical-
economic-conditions-for-injecting-hydrogen-into-natural-gas-ne.pdf
[25] MARCOGAZ. (2021). Odorisation of Natural Gas and Hydrogen Mixtures.
https://www.marcogaz.org/wp-content/uploads/2021/07/ODOR-Hydrogen-and-odorisation.pdf
[26] Tahir, M. Mohd., Ali, M. S., Salim, M. A., Bakar, R. A., Fudhail, A. M., Hassan, M. Z., &
Muhaimin, M. S. A. (2015). Performance Analysis of A Spark Ignition Engine Using Compressed
Natural Gas (CNG) as Fuel. Energy Procedia, 68, 355–362.
https://doi.org/10.1016/j.egypro.2015.03.266
[27] Menon, E. S. (2015). Transmission Pipeline Calculations and Simulations Manual (1st
edition). Gulf Professional Publishing.
[28] Khan, M. A., Young, C., MacKinnon, C., & Layzell, D. B. (2021). The Techno-Economics
of Hydrogen Compression. 1(1), 49.
[29] Witkowski, A., Rusin, A., Majkut, M., & Stolecka, K. (2018). Analysis of compression and
transport of the methane/hydrogen mixture in existing natural gas pipelines. International Journal
of Pressure Vessels and Piping, 166, 24–34. https://doi.org/10.1016/j.ijpvp.2018.08.002
[30] Lee, S.-W., Lee, H.-S., Park, Y.-J., & Cho, Y.-S. (2011). Combustion and emission
characteristics of HCNG in a constant volume chamber. Journal of Mechanical Science and
Technology, 25(2), 489–494. https://doi.org/10.1007/s12206-010-1231-5
[31] Verhelst, S., Demuynck, J., Sierens, R., Scarcelli, R., Matthias, N. S., & Wallner, T. (2013).
Update on the Progress of Hydrogen-Fueled Internal Combustion Engines. In Renewable
Hydrogen Technologies (pp. 381–400). Elsevier. https://doi.org/10.1016/B978-0-444-56352-
1.00016-7
[32] Tarkowski, R. (2019). Underground hydrogen storage: Characteristics and prospects.
Renewable and Sustainable Energy Reviews, 105, 86–94.
https://doi.org/10.1016/j.rser.2019.01.051
[33] Barbir, F., Basile, A., & Veziroğlu, T. N. (Eds.). (2016). Compendium of hydrogen energy.
Volume 3: Hydrogen energy conversion / edited by Frano Barbir, Angelo Basile and T. Nejat
Veziroglu. Elsevier ; WP Woodhead Publishing.
[34] Menon, E. S. (2005). Gas Pipeline Hydraulics (0 ed.). CRC Press.
https://doi.org/10.1201/9781420038224
[35] Demirel, Y. (2014). Nonequilibrium thermodynamics: Transport and rate processes in
physical, chemical and biological systems (Third edition). Boston ; Elsevier.

49
[36] Haeseldonckx, D., & Dhaeseleer, W. (2007). The use of the natural-gas pipeline infrastructure
for hydrogen transport in a changing market structure. International Journal of Hydrogen Energy,
32(10–11), 1381–1386. https://doi.org/10.1016/j.ijhydene.2006.10.018
[37] The American Society of Mechanical Engineers. (2020). ASME B31.8 Gas Transmission &
Distribution Piping Systems—ASME. New York, NY.
[38] The American Society of Mechanical Engineers. (2019). ASME B31.12 Hydrogen Piping &
Pipelines. New York, NY.
[39] Schmura, E., Klingenberg, M., & Corporation, C. T. (2005). Existing Natural Gas Pipeline
Materials and Associated Operational Characteristics.
[40] Laurent Bedel & Michel Junker. (2006). Natural gas pipelines for hydrogen transportation.
WHEC16: 16 World Hydrogen Energy Conference, France.
[41] Lee, J.A. (2016). NASA 2016-218602: Hydrogen Embrittlement. National Aeronautics and
Space Administration, Hunstville, AL.
[42] American Society for Testing and Materials. (2021). ASTM G129-21: Standard Practice for
Slow Strain Rate Testing to Evaluate the Susceptibility of Metallic Materials to Environmentally
Assisted Cracking. West Conshohocken, PA.
[43] American Petroleum Institute. (2016). API RP 941 Steels for Hydrogen Service at Elevated
Temperatures and Pressures in Petroleum Refineries and Petrochemical Plants. Washington D.C.
[44] Khare, A., Vishwakarma, M., & Ahmed, S. (2019). Combating Hydrogen Embrittlement with
Graphene Based Coatings. International Journal of Advanced Research In Engineering &
Technology, 10(6). https://doi.org/10.34218/IJARET.10.6.2019.027
[45] Hafsi, Z., Mishra, M., & Elaoud, S. (2018). Hydrogen embrittlement of steel pipelines during
transients. Procedia Structural Integrity, 13, 210–217. https://doi.org/10.1016/j.prostr.2018.12.035
[46] Shi, K., Xiao, S., Ruan, Q., Wu, H., Chen, G., Zhou, C., Jiang, S., Xi, K., He, M., & Chu, P.
K. (2022). Hydrogen permeation behavior and mechanism of multi-layered graphene coatings and
mitigation of hydrogen embrittlement of pipe steel. Applied Surface Science, 573, 151529.
https://doi.org/10.1016/j.apsusc.2021.151529
[47] Shi, K., Meng, X., Xiao, S., Chen, G., Wu, H., Zhou, C., Jiang, S., & Chu, P. K. (2021).
MXene Coatings: Novel Hydrogen Permeation Barriers for Pipe Steels. Nanomaterials, 11(10),
Article 10. https://doi.org/10.3390/nano11102737
[48] Li, X., Chen, L., Liu, H., Shi, C., Wang, D., Mi, Z., & Qiao, L. (2019). Prevention of Hydrogen
Damage Using MoS2 Coating on Iron Surface. Nanomaterials, 9(3), 382.
https://doi.org/10.3390/nano9030382
[49] NaturalHy. (2004). Preparing for the Hydrogen Economy by Using the Existing Natural Gas
System as a Catalyst. https://cordis.europa.eu/project/id/502661

50
[50] Musabikha, S., Utama, I. K. A. P., & Mukhtasor. (2018). State of the art in protection of
erosion-corrosion on vertical axis tidal current turbine. AIP Conference Proceedings, 1964(1),
020047. https://doi.org/10.1063/1.5038329
[51] Bahadori, A. (2015). Essentials of Coating, Painting, and Lining for the Oil, Gas and
Petrochemical Industries (1st edition). Gulf Professional Publishing.
[52] J, H., & Müller-Syring, G. (2006). Assessment of repair and rehabilitation technologies
relating to the transport of hythan (hydrogen-methane-mixture) (No. R0016-WP4). NaturalHY.
[53] Marina, T. & Markus, B. (2022). EUROPIPE Pipes internally lined with Epoxy Flow Coat
Ready for 100 Percent Hydrogen. Europipe. https://www.europipe.com/fileadmi n/europipe/ep-
docs/Hydrogen_Flow_Coat_January_2022_Final.pdf
[54] Ohaeri, E., Eduok, U., & Szpunar, J. (2018). Hydrogen related degradation in pipeline steel:
A review. International Journal of Hydrogen Energy, 43(31), 14584–14617.
https://doi.org/10.1016/j.ijhydene.2018.06.064
[55] Lauer, R. S. (2007, March 11). The Use Of High Performance Polymeric Coatings To Mitigate
Corrosion And Deposit Formation In Pipeline Applications. CORROSION 2007.
https://onepetro.org/NACECORR/proceedings-abstract/CORR07/All-CORR07/118400
[56] Najjar, Y. S. H. (2013). Hydrogen safety: The road toward green technology. International
Journal of Hydrogen Energy, 38(25), 10716–10728.
https://doi.org/10.1016/j.ijhydene.2013.05.126
[57] Gillette, J. L., & Kolpa, R. L. (2008). Overview of interstate hydrogen pipeline systems.
(ANL/EVS/TM/08-2). Argonne National Lab. (ANL), Argonne, IL (United States).
https://doi.org/10.2172/924391
[58] Rigas, F., & Sklavounos, S. (2005). Evaluation of hazards associated with hydrogen storage
facilities. International Journal of Hydrogen Energy, 30, 1501–1510.
https://doi.org/10.1016/j.ijhydene.2005.06.004
[59] Huising, O. J. C., & Krom, A. H. M. (2021, January 15). H2 in an Existing Natural Gas
Pipeline. 2020 13th International Pipeline Conference. https://doi.org/10.1115/IPC2020-9205
[60] HySafe. (2009). D113: Initial Guidance for Using Hydrogen in Confined Spaces—Results
from InsHyde. Sixth Framework Programme, Contract No SES6-CT-2004-502630.
http://www.hysafe.net/documents?deliverable=113
[61] Wahl, J., & Kallo, J. (2020). Quantitative valuation of hydrogen blending in European gas
grids and its impact on the combustion process of large-bore gas engines. International Journal of
Hydrogen Energy, 45(56), 32534–32546. https://doi.org/10.1016/j.ijhydene.2020.08.184
[62] Hormaza Mejia, A., Brouwer, J., & Mac Kinnon, M. (2020). Hydrogen leaks at the same rate
as natural gas in typical low-pressure gas infrastructure. International Journal of Hydrogen Energy,
45(15), 8810–8826. https://doi.org/10.1016/j.ijhydene.2019.12.159

51
[63] Leader, W., Partners, W., & Huld, T. (2001). EIHP2: Compilation of Existing Safety Data on
Hydrogen and Comparative Fuels. ENK6-CT2000-00442.
[64] Hübert, T., Boon-Brett, L., Black, G., & Banach, U. (2011). Hydrogen sensors – A review.
Sensors and Actuators B: Chemical, 157(2), 329–352. https://doi.org/10.1016/j.snb.2011.04.070
[65] Soundarrajan, P., & Schweighardt, F. (2008). Hydrogen Sensing and Detection. In R. Gupta
(Ed.), Hydrogen Fuel (pp. 495–534). CRC Press. https://doi.org/10.1201/9781420045772.ch15
[66] Geiger, G., Vogt, D., & Tetzner, R. (2006). State-of-the-Art in Leak Detection and
Localisation. Oil Gas European Magazine.
[67] Kida, T., Kuroiwa, T., Yuasa, M., Shimanoe, K., & Yamazoe, N. (2008). Study on the
response and recovery properties of semiconductor gas sensors using a high-speed gas-switching
system. Sensors and Actuators B: Chemical, 134(2), 928–933.
https://doi.org/10.1016/j.snb.2008.06.044
[68] Smith, J., Chae, J., Learn, S., Hugo, R., & Park, S. (2018, November 6). Pipeline Rupture
Detection Using Real-Time Transient Modelling and Convolutional Neural Networks. 2018 12th
International Pipeline Conference. https://doi.org/10.1115/IPC2018-78426
[69] Wylie, E. B. (1983). The Microcomputer and Pipeline Transients. Journal of Hydraulic
Engineering, 109(12), 1723–1739. https://doi.org/10.1061/(ASCE)0733-9429(1983)109:12(1723)
[70] Kunz, O., & Wagner, W. (2012). The GERG-2008 Wide-Range Equation of State for Natural
Gases and Other Mixtures: An Expansion of GERG-2004. Journal of Chemical & Engineering
Data, 57(11), 3032–3091. https://doi.org/10.1021/je300655b
[71] Hassanpouryouzband, A., Joonaki, E., Edlmann, K., Heinemann, N., & Yang, J. (2020).
Thermodynamic and transport properties of hydrogen containing streams. Scientific Data, 7(1),
222. https://doi.org/10.1038/s41597-020-0568-6
[72] Authors: Braga, A. S. (2018). Leakage Modeling Through Empirical Equations: An
Experimental Approach: (021). WDSA / CCWI Joint Conference Proceedings, 1.
https://ojs.library.queensu.ca/index.php/wdsa-ccw/article/view/12012
[73] Therkelsen, P., Werts, T., McDonell, V., & Samuelsen, S. (2009). Analysis of NOx Formation
in a Hydrogen-Fueled Gas Turbine Engine. Journal of Engineering for Gas Turbines and Power,
131(3). https://doi.org/10.1115/1.3028232
[74] T.M., S., Loughran, C.J., Jones, S., & Hopkins, E. (2009). API: Compendium of Greenhouse
Gas Emissions Methodologies for the Oil and Natural Gas Industry. American Petroleum Institute,
Washington D.C.
[75] Dokken, D. (2014). IPCC (Intergovernmental Panel on Climate Change): Impacts, Adaptation,
and Vulnerability. Cambridge University Press, Cambridge, UK.
[76] Dincer, I. (2012). Green methods for hydrogen production. International Journal of Hydrogen
Energy, 37(2), 1954–1971. https://doi.org/10.1016/j.ijhydene.2011.03.173

52
[77] Pinsky, R., Sabharwall, P., Hartvigsen, J., & O’Brien, J. (2020). Comparative review of
hydrogen production technologies for nuclear hybrid energy systems. Progress in Nuclear Energy,
123, 103317. https://doi.org/10.1016/j.pnucene.2020.103317
[78] Clean Energy Canada. (2020). A New Hope: How hydrogen can deliver climate solutions and
clean energy competitiveness for Canada. Vancouver, BC. https://cleanenergycanada.org/wp-
content/uploads/2020/10/CEC_Report_Hydrogen2020.pdf
[79] Bloomberg New Energy Finance (BNEF). (2021). Super-Cheap Hydrogen From Free Oil
Heralds Big Change: BNEF Q&A. https://proton.energy/wp-
content/uploads/2021/03/ProtonTechnologiesQA_Bloomberg.pdf
[80] Roy, J., & Demers, M. (2019). The hydrogen option for energy: A strategic advantage for
Quebec. Hydrogène Québec.
[81] International Renewable Energy Agency, IRENA. (2019). Hydrogen: A Renewable Energy
Perspective. 2nd Hydrogen Energy Ministerial Meeting, Tokyo, Japan.
https://www.irena.org/publications/2019/Sep/Hydrogen-A-renewable-energyperspective
[82] Li, J., Huang, H., & Kobayashi, N. (2017). Hydrogen combustion as a thermal source. Energy
Procedia, 142, 1083–1088. https://doi.org/10.1016/j.egypro.2017.12.360
[83] Altfeld, K., & Pinchbeck, D. (2013). Admissible hydrogen concentrations in natural gas
systems. Gas Energy.
[84] Choudhury, S., McDonell, V. G., & Samuelsen, S. (2020). Combustion performance of low-
NOx and conventional storage water heaters operated on hydrogen enriched natural gas.
International Journal of Hydrogen Energy, 45(3), 2405–2417.
https://doi.org/10.1016/j.ijhydene.2019.11.043
[85] Sun, M., Huang, X., Hu, Y., & Lyu, S. (2022). Effects on the performance of domestic gas
appliances operated on natural gas mixed with hydrogen. Energy, 244, 122557.
https://doi.org/10.1016/j.energy.2021.122557
[86] ISO/TC 197 Hydrogen technologies. (2019). ISO 14687:2019 Hydrogen fuel quality—
Product specification. Geneva, Switzerland.
[87] Society of Automotive Engineers (SAE). (2020). J2719: Hydrogen Fuel Quality for Fuel Cell
Vehicles. SAE International, Warrendale, PA.
[88] Du, Z., Liu, C., Zhai, J., Guo, X., Xiong, Y., Su, W., & He, G. (2021). A Review of Hydrogen
Purification Technologies for Fuel Cell Vehicles. Catalysts, 11(3), Article 3.
https://doi.org/10.3390/catal11030393
[89] Alqaheem, Y., Alomair, A., Mari, V., & Pérez, A. (2017). Polymeric Gas-Separation
Membranes for Petroleum Refining. International Journal of Polymer Science, 2017, 1–19.
https://doi.org/10.1155/2017/4250927

53
[90] Niemann, M. U., Srinivasan, S. S., Phani, A. R., Kumar, A., Goswami, D. Y., & Stefanakos,
E. K. (2008). Nanomaterials for Hydrogen Storage Applications: A Review. Journal of
Nanomaterials, 2008, e950967. https://doi.org/10.1155/2008/950967
[91] Office of Air and Radiation U.S. Environmental Protection Agency. (2008). Technical
Support Document for Hydrogen Production: Proposed Rule for Mandatory Reporting of
Greenhouse Gases.
[92] Cappellen, L. V., Croezen, H., & Rooijers, F. (2018). Feasibility Study into Blue Hydrogen:
Technical, Economic & Sustainability Analysis. CE Delft, Delft, The Netherlands.
https://cedelft.eu/publicaties/feasibility-study-into-blue-hydrogen/

54

You might also like