Next Article in Journal
An Optimized Power-Angle and Excitation Dual Loop Virtual Power System Stabilizer for Enhanced MMC-VSG Control and Low-Frequency Oscillation Suppression
Previous Article in Journal
Co-Design of a Wind–Hydrogen System: The Effect of Varying Wind Turbine Types on Techno-Economic Parameters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Systematic Literature Review on Pipeline Transport Losses of Hydrogen, Methane, and Their Mixture, Hythane

1
Department of Energy Engineering, Faculty of Energy Engineering and Industrial Management, University of Oradea, 410058 Oradea, Romania
2
Departament of Mechanical Engineering and Vehicles, Faculty of Management and Technological Engineering (ROMANIA), University of Oradea, 410058 Oradea, Romania
*
Author to whom correspondence should be addressed.
Energies 2024, 17(18), 4709; https://doi.org/10.3390/en17184709
Submission received: 19 August 2024 / Revised: 12 September 2024 / Accepted: 19 September 2024 / Published: 21 September 2024
(This article belongs to the Section A: Sustainable Energy)

Abstract

:
The transition to cleaner energy sources necessitates an in-depth understanding of the transport characteristics, losses, and opportunities associated with various gaseous fuels, including hydrogen, methane, and their mixtures, such as hythane. Hydrogen (H2), the most abundant element in the universe, is increasingly recognized as a viable alternative to fossil fuels, primarily due to its potential to reduce carbon footprints as a cleaner energy source. Gradually gaining prominence in the energy market, it is displacing other fuels such as methane. In some transport systems, hydrogen is mixed with methane (CH4) in order to reduce the carbon footprint while using the same existing production equipment. As more and more large methane consumers are implementing this mixture, we would like to see how the research has followed the market trend. An up-to-date research, development, and implementation status review is critical. This study aims to identify the main indicators of H2 and CH4 transport losses in pipes, providing a review of the state of the art in the specific literature. To deliver this, a systematic literature review (SLR) was carried out using preferred reporting items for systematic reviews and meta-analyses (PRISMA) methodology, pinpointing the research trends and results in peer review-published articles over a period of twelve years (2012–2024). Findings: this review identifies and points out, in numbers, the boundaries of the 2012–2024 timeline research.

1. Introduction

As global energy policies shift towards sustainability, the role of hydrogen and other gaseous fuels like methane in pipeline transportation becomes increasingly relevant. This review identifies the challenges and opportunities associated with the transport of hydrogen, methane, hythane (a blend of hydrogen and methane), and their mixtures. The aim is to synthesize existing knowledge and identify gaps for future research.
Transitioning from an energy market dominated by fossil fuels to one incorporating hydrogen offers numerous benefits, including environmental, social, economic, and energy-related advantages. Key stakeholders across political, environmental, and economic sectors are united in their support for the rapid implementation of this transition. The primary focus has been on hydrogen production and usage, but, similar to all energy forms and carriers, a big challenge lays in the transport and deployment from the point of production to the point of usage.
The development of new transport facilities for hydrogen (H2) is at odds with the need for infrastructure that can support the increasing demands of H2 production and consumption. A key strategic approach is to substitute methane (CH4) with H2, leveraging the substantial existing capital investments (CAPEX) that can be readily adapted to facilitate a blend of CH4 and H2, or even entirely pure H2. It is anticipated that the current CH4 transport infrastructure will play a crucial role in all strategic planning. Using the existing CH4 pipeline to transport hydrogen gas is a highly cost-effective solution for delivering significant quantities of H2.
In 2008, the only H2 pipe network was a limited one (750 km in USA and 1500 in EU [1]). In 2021, the USA had 2600 km [2], and the EU, in 2040, will have 40,000 km of hydrogen transport dedicated pipeline [3].
Utilizing CH4 infrastructure or incorporating H2 with CH4 up to a 30% volume of H2—commonly referred to as hythane—within the same transport system introduces potential unknown risks due to the differing chemical properties of H2 and CH4. These risks are not entirely understood or clearly defined and mitigated. It is essential to guarantee the same safety standards as those applied to conventional hydrocarbon technologies before implementing the transport network [4].
Blending hydrogen with natural gas can greatly cut greenhouse gas emissions, especially if the hydrogen comes from low-carbon sources like biomass, solar, wind, nuclear, or fossil fuels with carbon capture. The environmental benefits of sustainable hydrogen can be attributed to the natural gas blend, proportional to the hydrogen content. This is similar to the introduction of biogas into natural gas pipelines, offering a renewable option [5]. A credit trading system, akin to the renewable energy credit system in electricity, could incentivize the conversion of renewable energy into hydrogen, boosting the sustainability of natural gas [6,7].
In recent years, hydrogen and natural gas have gained recognition for their potential as clean energy fuels in the transportation sector, especially in ecological vehicles such as city buses. Hydrogen, in particular, offers the advantage of producing zero tailpipe emissions when used in fuel cells, making it a promising alternative to traditional fossil fuels. Similarly, natural gas, though a fossil fuel, produces fewer greenhouse gas emissions compared with gasoline or diesel. When used in a blended form, such as hythane (a mixture of hydrogen and natural gas), both fuels can contribute to reducing the overall carbon footprint of vehicles, while utilizing the existing fuel infrastructure. Several studies [8] have explored the use of hydrogen and natural gas in public transportation systems, highlighting their role in reducing emissions and improving air quality, particularly in urban areas. These alternative fuels represent a crucial step toward the decarbonization of the transportation sector, aligning with global sustainability goals.
When we are talking of H2 generation, we all see green H2, but the reality is that, as of 2021, 96% of world H2 generation is still made from fossil fuels and just 4% accounts for generation that could use only green energy [9,10]. Figure 1 shows the current potential share of renewable sources for hydrogen generation, along with the efficiency and costs associated with each generation method [10].
The generally recognized round-trip conversion efficiency of power to hydrogen and back to power is approximately 46% [10], so we must try to optimize all aspects of generation, storage, transport, and usage.
The generation of H2 being covered in another review [10] and storage being relatively simple, with little or no loss and with no higher CAPEX, we will try to focus on the transport side and the link between generation and usage.
While this review focuses on the transport systems for hydrogen, methane, and hythane, detailed topographical data for specific pipeline areas are unavailable. As a result, this study does not include topographical analysis. Future research may incorporate geographical variations to assess their impact on pipeline efficiency and pressure loss.
While the generation of hydrogen is addressed in other reviews [9] and storage is relatively straightforward with minimal loss and no significant increase in capital expenditure, our emphasis will be on the transportation aspect, which serves as the connection between generation and consumption.
In our paper, we provide a comprehensive analysis of the existing literature on the transport losses associated with hydrogen, methane, and hythane. Our contributions include the following:
  • Synthesis of current research. We systematically compile and synthesize research findings related to the losses experienced during the transportation of hydrogen, methane, and hythane, highlighting key factors that influence these losses.
  • Identification of knowledge gaps. Our review identifies critical gaps in the current body of knowledge, pointing out areas that require further investigation to optimize transport systems for these gases.
  • Comparative analysis. We offer a comparative assessment of transportation losses for hydrogen, methane, and hythane, providing insights into how the transport characteristics differ between these gases and the implications for efficiency and safety.
  • Recommendations for future research. Based on our findings, we propose recommendations for future research directions, focusing on enhancing transport efficiency and reducing losses, which can contribute to the development of more sustainable energy systems.
While significant advancements have been made in understanding the individual transport characteristics of methane and hydrogen, limited studies have systematically compared the transport losses of hydrogen, methane, and their mixture, hythane. This paper addresses this gap by providing a comprehensive analysis of transport losses across these gases, highlighting the implications for efficiency and safety in pipeline transport. By synthesizing research from various studies and providing a comparative framework, this review offers new insights that are critical for the development of future hydrogen–methane blended transport systems.
By addressing these aspects, our review not only consolidates existing knowledge but also serves as a foundation for future studies aimed at improving the transport systems for hydrogen, methane, and hythane.

2. Materials and Methods

Given the complexity and interdisciplinary nature of the research area that this study is focused on, a review is necessary to assess the field’s current readiness.
Identifying key features in peer-reviewed research papers—such as research focus, findings, economic factors, performance metrics, and technologies utilized—can serve as a foundational resource for manufacturers, researchers, managers, and other decision makers involved in H2 transport.
We use the systematic literature review (SLR) method as per preferred reporting items for systematic reviews and meta-analyses (PRISMA) [11] in order to deliver a review of the H2 transport topic accessible to researchers worldwide. This study follows a systematic literature review (SLR) methodology, which is designed to systematically collect, review, and analyze relevant research on the transport of hydrogen, methane, and their mixture, hythane. This method ensures that all research is comprehensively identified and evaluated. Furthermore, the review adheres to the preferred reporting items for systematic reviews and meta-analyses (PRISMA) guidelines, which provide a transparent and reproducible framework for conducting systematic reviews.
The objectives of this review are to identify the status of H2 transport research and market readiness.
The research questions raised during this review are as follows:
RQ1:
can a natural gas pipeline network be used for H2 transport?
RQ2:
what are the main pressure loss indicators in H2 transport?
RQ3:
are the risks of transporting H2 vs. CH4 higher?
In order to comply with the Kyoto Protocol [12] and the Paris Agreement [13], it is not enough to merely invest in hydrogen generation technologies and efficient usage, we also need to establish a transport network that can connect production and consumption at a finer resolution than what is currently available for methane (CH4).
This process follows three large steps [14,15]: (i) review the planning, (ii) conduct the review, and (iii) disseminate the results.
(i) During this stage, we define the review target and identify the limits of our research.
  • Pipes for H2 transport;
  • H2 and CH4 blended transport;
  • Replacement of CH4 with H2 in the actual pipe transport networks.
Based on our above targeted research area, we defined our procedure, like a guide that unites all the steps that are performed in this SLR. First, we defined our data pool, the EBSCO Discovery Service. This was our option because it contains an exhaustive list of research papers and because it has a no bias/high-quality policy.
Because a single search did not cover the whole spectrum of our research, we used the six strategies [5] defined in Table 1.
(ii) During the conduct the review stage, the database extracted from the EBSCO Discovery Service, namely 258 research papers, in July 2024 was considered.
(iii) The reporting and dissemination stage is covered under the results topic.
Due to the large number of research papers processed in this review, a flow of exclusion was identified and is presented in Figure 2. This flow chart follows the PRISMA methodology for reporting systematic reviews [11].

3. Results

3.1. RQ1: Can a Natural Gas Pipeline Network Be Used for H2 Transport?

Firstly, we need to examine the properties of both gases and determine how a functional structure designed to safely transport CH4 can meet the same functional and safety requirements for transporting H2.

3.1.1. Parametric Characteristics of the Two Gases

The physical proprieties and the chemical proprieties of CH4 are different from those of H2 (Appendix B). Also, the blend of the two gasses that are the scope of multiple investments WW is detailed in the extract hereunder (Table 2, fully reproduced in Appendix B [4,15,16,17,18,19,20,21,22]).
Hythane types are used both in automotive transport and in industry (30% H2), where the presence of H2 contributes to the engine combustion yield and to the CO2 emission reduction. Also, the mixture can be a competitive alternative fuel for existing combustion plants [18]. On the other hand, in discussions about the natural gas transportation network, there is typically an established upper limit of 10% hydrogen volume [23].
The properties of hydrogen (H2) and methane (CH4) (presented in Table 2) highlight significant differences that impact their transport through pipelines, both in existing infrastructure and future networks. The transition from methane (CH4) to hydrogen (H2) as an energy carrier presents several challenges, primarily due to the physical properties of hydrogen. Hydrogen’s lower molar mass and higher diffusivity allow it to permeate materials more easily than methane CH4, resulting in higher permeability in pipeline materials and a greater risk of embrittlement, particularly in steel pipes commonly used for methane transport. This compatibility issue is significant, as steel may become brittle and prone to cracking when exposed to hydrogen over time.
Hydrogen’s lower auto-ignition temperature and broader flammability limits compared with methane require stricter safety protocols to manage the associated risks in its transport and use. While methane is known for its higher calorific value by volume, which makes it more efficient for energy transport, blending hydrogen with methane—known as hythane—presents a potential solution that could strike a balance between lowering carbon emissions and leveraging the existing gas infrastructure [24].
However, adapting current pipeline systems to accommodate this H2/CH4 mixture poses challenges. Maintaining pipeline integrity and safety will necessitate significant modifications, especially for older networks that were not originally designed to accommodate the unique properties of hydrogen. These modifications may involve material upgrades, enhanced sealing techniques, and rigorous safety evaluations to ensure that both the efficiency of the energy transport and the safety of the infrastructure can be maintained [6].

3.1.2. H2 Transportation

We can see three main methods of H2 transportation.
  • In a gaseous state in specialized cylinders or tube trailers and pipeline transport (compressed hydrogen to high pressures—typically around 35–70 MPa—is one of the most common methods; this approach is suitable for short to medium distances and for smaller volumes);
  • In a liquid state (H2 can be transported in liquid form, which is advantageous for long distances and large-scale transport due to its higher energy density compared with gaseous hydrogen; however, the liquefaction process is energy-intensive and requires cryogenic temperatures around −253 °C);
  • In a bound form using solid or liquid carriers, including by land transport in cylindrical containers (H2 can also be transported using solid or liquid carriers, such as metal hydrides or chemical hydrogen storage systems [25]; these carriers allow for safer and more efficient storage and transportation, adding additional complexity to the hydrogen release process);
The safest and most practical method for transporting significant quantities of hydrogen (H2) over long distances is through pipelines in a compressed form. In order to achieve a technology readiness level of 9 and establish a cost-effective production method, hydrogen’s share in the energy market must account for at least 10% [26].
We will direct our focus on the existing natural gas transport piping network. Here, we can see that the vast majority are built out of steel [2,26]. Due to the size of H2 atoms, one proton and one electron, all metals are permeable to this gas [27], steel being one of them.
The H2 steel penetration from gas phase can be divided into the following stages [28].
  • Condensation of gaseous H2 on a steel surface (dispersive forces or Van der Waals forces) leads to surface coverage. The coverage of hydrogen atoms on the steel surface can be up to a monolayer under typical conditions. At room temperature and moderate pressures (0.1–1 MPa), surface coverage might range from 10⁻³ to 10⁻² molecules per square nanometer (molecules/nm²) [26,29].
  • Dissociation of molecules into atoms—chemisorption. The energy of chemisorption is the energy required for hydrogen molecules to dissociate and chemisorb onto steel surfaces and typically ranges between 20 and 50 kJ/mol. The degree of dissociation depends on temperature and surface properties but could result in a significant proportion of surface hydrogen in an atomic form at temperatures above 200 °C [26,29].
  • Transition of atoms through a steel surface—gas dissolution in steel. The surface permeation flux representing the flux of hydrogen atoms penetrating the steel surface (depending on pressure and temperature) could range between 10⁻⁶ and 10⁻⁴ mol/m²/s under typical industrial conditions (e.g., 200–400 °C and 1–10 MPa pressure). The actual flux depends significantly on the steel type and environmental conditions [26,29].
  • Diffusion of H2 atoms from the surface into the interior of the steel wall, characterized by:
    Diffusion coefficient. The diffusion coefficient of hydrogen in steel varies widely with temperature, typically ranging from 10⁻⁷ to 10⁻⁴ cm²/s at temperatures between 100 °C and 400 °C. At lower temperatures (e.g., 25 °C), the diffusion coefficient may be around 10⁻⁹ to 10⁻⁷ cm²/s [26,29].
    Penetration depth. The penetration depth of hydrogen into steel over extended periods (e.g., 1000 h) at room temperature might be in the range of 10 to 100 micrometers (µm), increasing significantly at higher temperatures due to the exponential increase in the diffusion coefficient [26,29].
The hydrogenation intensity is determined by the gaseous H2 pressure, temperature, and type of steel. The solubility of hydrogen in iron, which indicates the maximum amount of hydrogen that can exist in a solid solution form in Fe at a temperature of 23 °C and a pressure of 0.1 MPa, is ~0.0001 m3/0.1 kg Fe [26]. This process is directly proportional to the working pressure and the strength of the steel. The hydrogenation of steel pipes changes the mechanical properties of steel pipes and increases the risk of failure in connecting surfaces (welding, T junctions, etc.) [2,26,30].
Some issues related to the injection of H2 into the CH4 pipes are related to the pressure difference between the two gasses. This pressure difference leads to a Joule–Thomson effect, causing the temperature of the H2 to increase during its expansion (pressure decrease) [4].
Having such a high number of increased risks for transportation in existing natural gas steel pipes, we would assume that a steel piping network would not be a viable solution for H2 transport. However, mitigating the risks (as performed in early CH4 pipe systems) the economical and sustainability criteria are the ones that drive the transport industry [31].
Based on the status outlined, it appears that the strategy involves repurposing the existing steel network instead of implementing significant changes. This approach leverages the current infrastructure while adapting to a new risk matrix that addresses any potential challenges. By doing so, the system can maintain the same safety standards that are currently in place for methane usage. This strategy not only ensures continuity of operations but also optimizes the existing resources in a way that aligns with safety and risk management objectives [27].

3.2. RQ2: What Are the Main Pressure Loss Indicators in H2 Transport?

The main research indicators identified in this review are as follows:
  • Mixture H2/CH4 [%];
  • Gas or liquid state;
  • Pipeline diameter [mm];
  • Wall thickness [mm];
  • Length of pipe [m];
  • Pipe material;
  • Strength class;
  • Outside pipe pressure [Pa];
  • Inside pipe pressure [Pa];
  • Velocity [m/s];
  • Temperature [°C];
  • Diffusion coefficient of H2 in metal [m2/s];
  • Solubility of H2 in Fe [m3/m3 solid*MPa0.5];
  • H2 leaks [m3];
  • Power usage for transport [MW];
  • Load loss [MPa];
  • Inner surface roughness of pipe [μm];
  • Loss due to diffusion [%]
  • Enthalpic jump [kJ/kg];
  • Mass flow [kg/s];
  • Pipe production technology.
In order to visualize the focus of the research on hythane transport, we present hereunder some charts to summarize the KPIs and results. Figure 3 presents the specified working pressure of the hythane (hydrogen–methane) transported [25,26,27,28,29,30,31,32,33,34].
Most research (94.12%) is aimed at the gas transport of H2 and only 5.88% is directed to the liquid phase of H2 [29,35]. This preference for the gas phase of hythane and H2 is determined by the economics of reusing or sharing the existing infrastructure of CH4 transport. Figure 4 displays a box-and-whisker plot illustrating the pipe parameters of the modeled or physically studied pipeline network. In Figure 5, we add the materials used in the research regarding the hythane transport in real or modeled networks [35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54].
The coating of pipelines [55] or new methods (e.g., pulsed laser ablation [44]) of delivering pipes for hythane or H2 transport could reduce the associated risks. Building pipelines for H2 transport only drastically differs in costs. Estimated costs of operating the existing CH4 pipeline infrastructure for H2 or hythane increases the cost by 30% to 50% more than CH4 [24].
We identified only one paper that mentions two indicators related to the velocity increase and pressure drop after injecting/blending H2 into CH4 [56]. For 17% H2 in hythane, we see a velocity increase of 0.16 [m/s] and a pressure drop of 0.019 [MPa].
In Figure 6, we highlight the outline of the transport pressure, velocity, and temperature (only for the gas phase of H2; for the liquid phase, we have only one research at −252 °C [41]).
As most of our research focus includes steel pipes, we mention some characteristics that define the transport and material specifications.
  • Solubility of H2 in iron 0.0001 cm/0.1 kg Fe [26].
  • Solubility of H2 in steel 5.9 × 10−5 [m3 gas/(m3 solid*MPa0.5)] [26].
  • Diffusion coefficient of H2 in steel [m2/s] varies from
    iron—5.8 × 10−10 [28];
    K42 steel—4.32 × 10−11 [26];
    100Cr6 steel—4.9 × 10−12 [42];
    304 steel—7.37 × 10−16 [28].
  • H2 permeability of the piping coating:
    -
    Diglycidyl ether (DGEBA)/polyetheramine (D-400) epoxy coating, 0.35 Barrer [55];
    -
    Crosslinked poly (vinyl alcohol) coating, 0.084 Barrer [55].

3.3. RQ3: Are Risks of Transporting H2 vs. CH4 Higher?

In order to identify and assess the risk associated with each process and segment of the transportation area, a standard risk matrix is provided in Table 3. The matrix is suitable for CH4 and for H2 [57,58].
In Table 4, we try to summarize the upper limit of hythane H2% blending so that the risk remains a green one (as per above matrix) [59].
In Figure 7, we present the maximum acceptable H2 mixture in CH4 transport within piping networks across different countries. Any higher mixture of H2 into CH4 would increase the risk associated with yellow labeling [56,57,60].
Due to the proprieties of H2 versus CH4 mentioned in Table 2, the risk assessment for the same event could be higher for H2 on a severity scale and on occurrence probability even though the associated hazards are similar [61,62]. All the assessment vaguely presented in Table 3 should take into consideration the following:
  • Correctly assessing the vulnerabilities of metal piping;
  • Understanding the behavior of H2 released into the atmosphere, including its rapid dispersal and tendency not to accumulate near the ground, where ignition sources are more likely to be present;
  • Accidental release of H2 means a much higher flammable mixture, necessitating a review of applicable safety measures;
  • Compared with methane (CH4), hydrogen possesses a significantly higher flame propagation velocity, which can increase the risk of fire hazards;
  • The flammability and radiation hazard should be reconsidered;
  • Equipment built for use in gas group IIC (ATEX) is to be used [58];
  • Gas measurement and leak detection equipment must be specifically suitable for hydrogen to effectively identify potential leaks and mitigate risks associated with its release.

4. Expanded Comparative Analysis and Safety Considerations in Hydrogen and Methane Transport Systems

Although hydrogen and methane share many similarities, there are notable differences that lead to a knowledge gap, as presented in Table 5, primarily resulting from the insufficient experimental data, which have largely focused on decades of methane transport. In comparing methane and hydrogen for transport and safety considerations, it is essential to examine their respective physical and chemical properties. These properties directly influence their behavior during pipeline transport and the associated risks, such as energy losses, safety hazards, and transport efficiency. Table 5 highlights the key differences between methane and hydrogen, focusing on properties relevant to their transport and safety performance.
RQ1: can the natural gas pipeline network be used for H2 transport?
The existing natural gas (CH4) pipeline infrastructure has the potential to be repurposed for hydrogen (H2) transport, but several technical challenges and limitations need to be addressed. Due to its status as the smallest molecule, hydrogen exhibits greater diffusivity and permeability than methane. These properties increase the likelihood of leakage and raise the risk of embrittlement in metals, particularly in the high-strength steel typically utilized in natural gas pipelines [58,59].
Hydrogen, methane, and their mixture, hythane, exhibit distinct transport losses due to their differing physical and chemical properties. Hydrogen, with its lower molecular weight and higher diffusivity, tends to leak more easily from pipelines, leading to greater pressure losses compared with methane. On the other hand, methane, being denser, experiences lower leakage rates but higher energy losses over long transport distances due to friction and viscosity. Hythane, as a blend of hydrogen and methane, balances these properties, but the exact losses depend on the ratio of hydrogen to methane in the mixture. This review highlights that the efficiency of transport systems for hydrogen will require significant adaptations in pipeline materials and sealing technologies to minimize losses, whereas methane can be transported more reliably using the current infrastructure.
Studies have shown that blending up to 20% H2 by volume with natural gas can be managed within the existing infrastructure with minimal modifications [24]. However, as the concentration of hydrogen increases, so do the risks of pipeline embrittlement, leakage, and reduced integrity of the pipeline materials [33]. In recent years, several real-world projects have tested the feasibility of hydrogen transport through existing natural gas pipelines. For instance, the “HyDeploy” project in the United Kingdom demonstrated that up to 20% hydrogen could be blended into the existing natural gas network without significant safety or efficiency issues [63]. However, this study also noted challenges related to pipeline materials, particularly the need for coatings and seals that are resistant to hydrogen embrittlement [63]. Similarly, in the United States, the “H2@Scale” initiative has identified key technical barriers, including pressure regulation and monitoring for hydrogen leakage [64]. These examples underline the importance of tailored solutions for hydrogen transport, which must account for both material compatibility and operational safety. Research indicates that additional measures, such as the application of internal coatings, the use of low-carbon steel, and ongoing monitoring of pipeline integrity, would be necessary to mitigate these risks [29,34,35].
Moreover, the operational conditions for H2 transport differ from those for CH4. Hydrogen’s lower energy density requires higher pressures to achieve the same energy throughput as methane [65]. This difference could necessitate modifications in compressors and other pipeline network components.
Several studies have conducted comparative analyses on the transport efficiency and safety of hydrogen versus methane. For example, a recent study [66] found that the energy loss per kilometer of pipeline is approximately 15% higher for hydrogen than for methane under similar conditions, primarily due to hydrogen’s higher diffusivity. Another comparative study [67] demonstrated that, while hydrogen offers a cleaner alternative, the cost of retrofitting pipelines to safely handle hydrogen adds significant operational expenses. In contrast, methane, while less prone to leakage, poses greater environmental risks due to its higher greenhouse gas potential. These findings highlight the trade-offs involved in transitioning from methane to hydrogen in existing pipeline networks. Pipelines designed for natural gas or hythane do not face the same degree of embrittlement or permeation challenges. However, for the safe transport of hydrogen, solutions such as composite materials, lined pipelines, or advanced coatings are essential. These materials not only provide the necessary mechanical strength but also prevent hydrogen from penetrating the pipeline structure, reducing the risk of leakage and maintaining pipeline integrity over time [68]. Our review emphasizes the need for more robust material standards and suggests future research on developing cost-effective materials that can accommodate both hydrogen and methane in blended systems [69]. Recent advancements in materials science have enabled the development of high-strength polymers and composites that can withstand the higher pressures required for hydrogen transport [70], while also preventing hydrogen diffusion through the pipeline walls. Compared with traditional steel, these flexible systems not only reduce the risk of pipeline failure but also allow for easier installation and maintenance. Additionally, the use of flexible pipelines in hydrogen distribution networks provides a cost-effective alternative to retrofitting existing natural gas pipelines
RQ2: what are the main pressure loss indicators of H2 transport?
Pressure losses in hydrogen transport are significantly influenced by the unique physical properties of hydrogen, which differ from those of methane. These properties include lower density and higher diffusivity, both of which contribute to greater pressure loss over long pipeline distances. In particular, hydrogen’s high flow velocity at the same pressure gradient as methane leads to greater frictional resistance in pipelines. Studies have shown that maintaining optimal flow velocity is crucial to minimizing these losses. Furthermore, the inner surface roughness of pipelines, which affects friction, plays a larger role in hydrogen transport. The use of smoother internal surfaces in hydrogen pipelines may reduce frictional losses and help maintain pressure over long distances.
Pressure loss in hydrogen pipelines is influenced by several factors, which differ from those in natural gas pipelines due to hydrogen’s unique properties. The primary indicators of pressure loss in hydrogen transport include the following:
  • Pipe diameter and wall thickness. Smaller pipe diameters and thicker walls generally result in greater pressure losses due to increased friction and reduced flow area. For hydrogen, which is less dense and more prone to turbulent flow, optimizing pipe dimensions is crucial to minimize losses.
Fundamental relationship is the relationship between pipe diameter D, flow rate Q, and pressure drop ΔP in a pipeline carrying hydrogen, described by the Darcy–Weisbach Equation (1):
P =   f L D   ρ v 2 2
where f is the friction factor (which depends on the flow regime and pipe roughness), L is the length of the pipe, D is the pipe diameter, ρ is the density of the hydrogen, and v is the average flow velocity of the hydrogen.
Flow rate relationship. The flow rate Q is related to the velocity v and the cross-sectional area A of the pipe by:
Q = A v = π   D 2 4 v
Substituting into the Darcy–Weisbach equation illustrates that for, a given flow rate Q, increasing the pipe diameter D reduces the velocity v and, therefore, the pressure drop ΔP.
  • Average flow velocity. Hydrogen has a higher flow velocity at the same pressure gradient compared with methane. This increased velocity can lead to higher frictional losses, especially in long-distance transport pipelines. Maintaining optimal flow velocity is essential to reducing pressure drop.
  • Surface roughness. The internal surface roughness of the pipeline has a significant impact on pressure loss. Hydrogen’s low viscosity exacerbates the effects of surface roughness, increasing frictional resistance. Pipelines designed for hydrogen transport might require smoother internal surfaces to mitigate these losses [71].
  • Operating pressure and temperature. Higher operating pressures tend to reduce pressure losses due to a higher density of hydrogen, which can compensate for its low molecular weight. However, operating at elevated temperatures can decrease hydrogen’s density, increasing pressure loss.
Pressure losses in hydrogen transport are significantly influenced by the unique physical properties of hydrogen, which differ from those of methane [72]. These properties include lower density and higher diffusivity, both of which contribute to greater pressure loss over long pipeline distances. In particular, hydrogen’s high flow velocity at the same pressure gradient as methane leads to greater frictional resistance in pipelines. Studies [72,73] have shown that maintaining optimal flow velocity is crucial to minimizing these losses. Furthermore, the inner surface roughness of pipelines, which affects friction, plays a larger role in hydrogen transport. The use of smoother internal surfaces in hydrogen pipelines may reduce frictional losses and help maintain pressure over long distances.
RQ3: are the risks of transporting H2 higher than CH4?
The transportation of hydrogen poses several distinct risks compared with methane. One of the primary concerns is hydrogen embrittlement, a phenomenon where hydrogen atoms penetrate steel pipelines, causing the material to become brittle and more prone to cracking. This issue is exacerbated in high-strength steel pipelines traditionally used for methane transport. Hydrogen leakage is also a more significant risk due to its smaller molecular size, which allows it to escape through pipeline joints and even through the steel itself. Furthermore, hydrogen’s broader flammability limits and lower ignition energy increase the risk of accidental ignition in the event of a leak. These factors necessitate stricter safety protocols, including the use of advanced coatings and materials that can withstand hydrogen’s unique properties.
The primary risks include the following:
  • Hydrogen embrittlement. Hydrogen can penetrate steel pipelines, leading to embrittlement—a phenomenon whereby the metal becomes brittle and more susceptible to cracking. This risk is particularly pronounced in high-strength steels and requires the use of special alloys or coatings to mitigate [28,29,42].
  • Leakage. Due to its small molecular size, hydrogen is more likely to leak through pipeline joints, fittings, and even through the steel itself. This increases the risk of explosions, especially since hydrogen has a wide flammability range (4–75% in air) and a low ignition energy compared with methane [43,57,74].
  • Explosion hazard. Hydrogen’s high diffusivity and wide flammability limits make it more prone to accidental ignition. When leaks occur, hydrogen can form explosive mixtures more readily than methane, especially in confined spaces. The energy released in hydrogen explosions is also typically higher, posing a significant safety risk [7].
  • Joule–Thomson effect. When hydrogen is released from high pressure (as might happen during a leak), its temperature increases, unlike methane, which cools upon expansion. This temperature rise can lead to ignition risks if there are sparks or other ignition sources nearby [36].
  • Accidental damage to a pipe in an excavation activity. Here, the failure frequencies are similar to those for natural gas “from 2016 to 2022 it resulted in 11 fatalities, 36 injuries, and over USD 144 M in damages” [74] in the US only.
Figure 8, hereunder, provides a comparative timeline of major accidents associated with methane (CH4) and hydrogen (H2) transport systems. The upper chart illustrates significant incidents involving CH4, highlighting the frequency and severity of accidents over the years. The lower chart focuses on hydrogen-related incidents, underscoring the unique risks posed by hydrogen transport due to its higher diffusivity, lower ignition energy, and broader flammability limits compared with methane. This timeline emphasizes the importance of understanding the distinct safety challenges associated with hydrogen transport as the energy sector transitions towards cleaner alternatives.

5. Conclusions

The transition from methane (CH4) to hydrogen (H2) as an energy carrier presents both opportunities and challenges in the context of existing transport infrastructure. This analysis outlines key strategies and considerations for achieving this shift effectively while ensuring safety and efficiency.
  • Incremental integration of hydrogen. The initial phase of integrating hydrogen into existing natural gas systems involves blending H2 with methane in concentrations ranging from 5% to 30%. This blending is contingent upon local regulations and, as such, the adjustment of risk assessment matrices and maintenance practices for metal piping and associated infrastructure becomes imperative. Effective management of this transition requires a thorough understanding of the materials properties and behavior of piping systems under mixed gas conditions.
  • Adaptation of combustion systems. Converting large-scale CH4 combustion systems to utilize hythane (a mixture of hydrogen and natural gas) up to 30% H2 allows for a competitive alternative that aids in maintaining the necessary flow within transport pipelines. This conversion not only facilitates the transition to hydrogen but also leverages existing infrastructure and technologies for a more gradual shift.
  • Innovative liquid hydrogen transport. Exploring the feasibility of transporting liquid hydrogen through cryogenic pipelines presents a revolutionary approach to reducing transport costs, with theoretical reductions in energy costs of up to 99.5% compared with traditional methane transport [41]. This method could potentially open new avenues for hydrogen distribution, although it requires innovative material solutions to handle the extreme conditions.
  • Repurposing natural gas pipelines. While repurposing existing natural gas pipelines for hydrogen transport is achievable, it necessitates extensive upgrades to ensure safety and efficiency. This includes material adjustments to withstand hydrogen’s unique properties, the implementation of enhanced monitoring systems, and limitations on hydrogen concentration to mitigate mechanical stress risks. Additionally, integrating an inner lining designed to reduce friction could further optimize energy loss during transport [55].
  • Advanced pipeline design. The introduction of pipes with a 3D-structured surface—such as triangular or knife-blade geometries—aimed at minimizing drag, represents a cutting-edge solution for enhancing hydrogen flow efficiency. These advanced designs can significantly reduce turbulence and pressure losses, thus optimizing the overall performance of hydrogen transport systems.
  • Optimization of transport parameters. To ensure efficient hydrogen transportation, a comprehensive analysis of pipeline dimensions, flow velocity, surface roughness, and operating conditions is critical. Effective design and optimization of these factors are essential to minimize pressure losses and maximize throughput.
  • Risk management. The inherent risks associated with hydrogen transport—such as embrittlement of materials, leakage, and explosion potential—are notably higher than those for methane. Addressing these risks demands substantial advancements in technology and materials science, along with rigorous safety protocols that may include continuous monitoring and rapid response mechanisms.
In conclusion, the transition to a hydrogen economy is fraught with both challenges and potential. By strategically adapting infrastructure, investing in innovative transport methods, and rigorously managing safety risks, it is possible to pave the way for a more sustainable and efficient hydrogen transport ecosystem. The steps outlined provide a foundational framework for stakeholders in the energy sector as they navigate this critical evolution.
Research on hydrogen and hythane (a blend of hydrogen and natural gas) transport losses is critical for improving the efficiency and sustainability of these energy carriers. A few ideas for future research focused on the following topics:
  • Leak detection technologies. Development of advanced sensors and monitoring systems for real-time detection of hydrogen and hythane leaks in pipelines. This can include research on nanomaterials or IoT-enabled devices for enhanced sensitivity.
  • Material innovation. Investigating new materials for piping and storage that reduce permeability to hydrogen and minimize transport losses. This could include coatings or composite materials that are more resistant to hydrogen embrittlement.
  • Integration with renewable energy sources. Analyzing how integrating hydrogen transport with renewable energy systems (like solar or wind) can enhance energy efficiency and minimize loss during transport.
By pursuing these research avenues, experts can help optimize hydrogen and hythane transport systems, greatly improving their feasibility and efficiency as future energy sources.
This paper provides a novel contribution to the existing body of research by systematically reviewing and comparing the transport losses of hydrogen, methane, and hythane. Our findings highlight the differences in transport characteristics between these gases and identify the technical and safety challenges associated with hydrogen transport. By filling the current gaps in the literature, this review serves as a foundation for future studies aimed at optimizing transport efficiency and enhancing the safety of pipeline systems. The comparative analysis presented here will help industry stakeholders make informed decisions regarding the integration of hydrogen into existing methane transport infrastructures

Author Contributions

Conceptualization, F.C.D. and C.H.; methodology, F.C.D., D.-C.S. and H.N.H.; validation, D.-C.S.; formal analysis, C.H., D.-C.S. and H.N.H.; investigation, F.C.D.; resources, C.H.; data curation, F.C.D.; writing—original draft preparation, F.C.D.; visualization, C.H., D.-C.S. and H.N.H.; supervision, C.H.; funding acquisition, C.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by University of Oradea, within the Grants Competition “Scientific Research of Excellence Related to Priority Areas with Capitalization through Technology Transfer: INO-TRANSFER-UO”, Project No. 315/21/12/2021; “Scientific Research of Excellence Related to Priority Areas with Capitalization through Technology Transfer: INO-TRANSFER-UO—2nd Edition”, Project No. 239/01/11/2022; and the APC was funded by University of Oradea.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
3DThree-Dimensional
BWBox and Whiskers Chart
CAPEXCapital Expenditures
FeIron
H2Hydrogen
KPIKey Performance Indicator
NAFNot Accounted For
OPEXOperating Expenses
PRISMAPreferred Reporting Items for Systematic Reviews and Meta-Analyses
RQResearch Questions
SLRSystematic Literature Review
WWWorld Wide

Appendix A

1.
Disciplines used for each criterion: Agriculture and Agribusiness, Applied Sciences, Architecture, Biotechnology, Chemistry, Computer science, Construction and Building, Earth and Atmospheric sciences, Engineering, Environmental Sciences, Life Sciences, Mining and natural resources, Power and Energy, Science, Technology
2.
Search mode used for criterion 1 to 5:
Find all my search terms;
Also search within the full text of the article;
Apply equivalent subjects.
3.
Search mode used for criterion 6: online recommended research papers during skim and scan.

Appendix B

Physical proprieties of H2, CH4, and mixtures (volume %).
PropertiesHydrogen (H2)Methane (CH4)Hythane
10% H2-90% CH4
Hythane
25% H2-75% CH4
Hythane
30% H2-70% CH4
Mixture 40% H2-60% CH4Mixture 60% H2-40% CH4Unit
Molar mass2.0216.0414.6411.4815.0610.437.63g/mole
Critical temperature33.20190.65174.91139.48179.63127.6796.18C
Critical pressure1.3154.5404.218-4.314--MPa
Vapor density at normal boiling point1.341.821.77-1.79--kg/m3
Vapor density at 20 C and 0.1 MPa0.08380.65100.59430.46670.61130.42410.3107kg/m3
Specific heat capacity at 20 C and constant pressure14.402.213.436.173.067.099.52kJ/kg/K
Specific heat ratio (Cp/Cv)1.401.311.321.341.321.351.36
Lower calorific value by mass (lower heating value, weight basis)120.0048.0055.2071.4053.0476.8091.20MJ/kg
Lower calorific value by volume at 1 atm11.0035.0032.6027.2033.3225.4020.60MJ/m3
Higher calorific value by mass142.0053.0061.9081.9359.2388.60106.40MJ/kg
Higher calorific value by volume at 1 Atm13.0039.0036.0030.0027.0028.0020.00MJ/m3
Maximum flame temperature1526.851221.851252.351320.981243.21343.851404.85°C
Explosive (deniability) limits18.205.70---10.7013.20Vol % in air
Limiting oxygen for combustion5.0012.00-----Vol %
Flammability limits4.105.304.404.404.504.604.60Vol % in air
Auto-ignition temperature560.00600.00590.00590.00580.00580.00570.00°C
Laminar burning velocity3.100.400.681.310.601.522.08m/s
Dilute gas viscosity at T ¼ 299 K0.0000090.0000110.0000110.0000100.0000110.0000100.000010Pa × s
Molecular diffusivity in air0.0000610.0001600.0001500.0001280.0001530.0001200.000101m2/s
Solubility in water0.00160.02500.02270.01740.02340.01560.0110kg/m3

References

  1. Grasso, N. Fire prevention technical rule for gaseous hydrogen transport in pipelines. Int. J. Hydrogen Energy 2008, 34, 4675–4683. [Google Scholar] [CrossRef]
  2. Paul, W. Pipeline Transportation of Hydrogen: Regulation, Research, and Policy. Congressional Research Service. Available online: https://www.everycrsreport.com/files/2021-03-02_R46700_294547743ff4516b1d562f7c4dae166186f1833e.pdf (accessed on 30 September 2022).
  3. Europe Could Operate 40,000 km of Hydrogen Pipelines by 2040. Reuters 2022. Available online: https://www.reuters.com/business/sustainable-business/europe-could-operate-40000-km-hydrogen-pipelines-by-2040-operators-2021-04-13/ (accessed on 12 August 2024).
  4. Messaoudani, Z.L.; Rigas, F.; Hamid, M.D.B.; Hassan, C.R.C. Hazards, safety and knowledge gaps on hydrogen transmission via natural gas grid: A critical review. Int. J. Hydrogen Energy 2016, 41, 17511–17525. [Google Scholar] [CrossRef]
  5. Lee, H.; Lee, S. Economic Analysis on Hydrogen Pipeline Infrastructure Establishment Scenarios: Case Study of South Korea. Energies 2022, 15, 6824. [Google Scholar] [CrossRef]
  6. Melaina, M.W.; Antonia, O.; Penev, M. Blending Hydrogen into Natural Gas Pipeline Networks: A Review of Key Issues. NREL. Available online: http://www.nrel.gov/docs/fy13osti/51995.pdf (accessed on 15 November 2022).
  7. Vidas, L.; Castro, R.; Pires, A. A Review of the Impact of Hydrogen Integration in Natural Gas Distribution Networks and Electric Smart Grids. Energies 2022, 15, 3160. [Google Scholar] [CrossRef]
  8. Pyza, D.; Gołda, P.; Sendek-Matysiak, E. Use of hydrogen in public transport systems. J. Clean. Prod. 2022, 335, 130247. [Google Scholar] [CrossRef]
  9. Zhang, B.; Zhang, S.-X.; Yao, R.; Wu, Y.-H.; Qiu, J.-S. Progress and prospects of hydrogen production: Opportunities and challenges. J. Electron. Sci. Technol. 2021, 19, 100080. [Google Scholar] [CrossRef]
  10. Hora, C.; Dan, F.C.; Rancov, N.; Badea, G.E.; Secui, C. Main Trends and Research Directions in Hydrogen Generation Using Low Temperature Electrolysis: A Systematic Literature Review. Energies 2022, 15, 6076. [Google Scholar] [CrossRef]
  11. Page, M.J.; Moher, D.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E. PRISMA 2020 explanation and elaboration: Updated guidance and exemplars for reporting systematic reviews. BMJ 2021, 372, n160. [Google Scholar] [CrossRef]
  12. Corporate Author, Kyoto Protocol. 2017. Available online: http://unfccc.int/kyoto_protocol/items/2830.php (accessed on 12 August 2024).
  13. Paris Agreement to the United Nations Framework Convention on Climate Change. 2015. Available online: https://unfccc.int/sites/default/files/english_paris_agreement.pdf (accessed on 12 August 2024).
  14. Pires, A.L.G.; Junior, P.R.; Morioka, S.N.; Rocha, L.C.S.; Bolis, I. Main Trends and Criteria Adopted in Economic Feasibility Studies of Offshore Wind Energy: A Systematic Literature Review. Energies 2021, 15, 12. [Google Scholar] [CrossRef]
  15. Kovács, K.E.; Dan, B.; Hrabéczy, A.; Bacskai, K.; Pusztai, G. Is Resilience a Trait or a Result of Parental Involvement? The Results of a Systematic Literature Review. Educ. Sci. 2022, 12, 372. [Google Scholar] [CrossRef]
  16. Ciccarelli, G.; Chaumeix, N.; Mendiburu, A.Z.; N’Guessan, K.; Comandini, A. Fast-flame limit for hydrogen/methane-air mixtures. Proc. Combust. Inst. 2019, 37, 3661–3668. [Google Scholar] [CrossRef]
  17. Scientific Data Curation Team. Thermodynamic and transport properties of hydrogen containing streams. Sci. Data 2020, 7, 222. [Google Scholar] [CrossRef] [PubMed]
  18. Ilbas, M.; Crayford, A.; Yilmaz, I.; Bowen, P.; Syred, N. Laminar-burning velocities of hydrogen–air and hydrogen–methane–air mixtures: An experimental study. Int. J. Hydrogen Energy 2006, 31, 1768–1779. [Google Scholar] [CrossRef]
  19. Cheng, R.K.; Oppenheim, A.K. Autoignition in methane-hydrogen mixtures. Combust. Flame 1984, 58, 125–139. [Google Scholar] [CrossRef]
  20. Conti, R.S.; Hertzberg, M. Thermal Autoignition Temperatures for Hydrogen-Air and Methane-Air Mixtures. J. Fire Sci. 1988, 6, 348–355. [Google Scholar] [CrossRef]
  21. Miao, H.; Lu, L.; Huang, Z. Flammability limits of hydrogen-enriched natural gas. Int. J. Hydrogen Energy 2011, 36, 6937–6947. [Google Scholar] [CrossRef]
  22. Vandenschoor, F.; Verplaetsen, F. The upper flammability limit of methane/hydrogen/air mixtures at elevated pressures and temperatures. Int. J. Hydrogen Energy 2007, 32, 2548–2552. [Google Scholar] [CrossRef]
  23. Pandey, A.; Mohan, S.V.; Chang, J.-S.; Hallenbeck, P.C.; Larroche, C. (Eds.) Biohydrogen. In Biomass, Biofuels, Biochemicals, 2nd ed.; Elsevier: Amsterdam, The Netherlands; Cambridge, MA, USA, 2019. [Google Scholar]
  24. Mahajan, D.; Tan, K.; Venkatesh, T.; Kileti, P.; Clayton, C.R. Hydrogen Blending in Gas Pipeline Networks—A Review. Energies 2022, 15, 3582. [Google Scholar] [CrossRef]
  25. Veluswamy, H.P. Energy Storage in Hydrates: Status, Recent Trends, and Future Prospects. ACS Appl. Energy Mater. 2024. [Google Scholar] [CrossRef]
  26. Bolobov, V.I.; Latipov, I.U.; Popov, G.G.; Buslaev, G.V.; Martynenko, Y.V. Estimation of the Influence of Compressed Hydrogen on the Mechanical Properties of Pipeline Steels. Energies 2021, 14, 6085. [Google Scholar] [CrossRef]
  27. Schefer, R.; Houf, W.; Sanmarchi, C.; Chernicoff, W.; Englom, L. Characterization of leaks from compressed hydrogen dispensing systems and related components. Int. J. Hydrogen Energy 2006, 31, 1247–1260. [Google Scholar] [CrossRef]
  28. Li, X.; Ma, X.; Zhang, J.; Akiyama, E.; Wang, Y.; Song, X. Review of Hydrogen Embrittlement in Metals: Hydrogen Diffusion, Hydrogen Characterization, Hydrogen Embrittlement Mechanism and Prevention. Acta Metall. Sin. (Engl. Lett.) 2020, 33, 759–773. [Google Scholar] [CrossRef]
  29. Li, H.; Niu, R.; Li, W.; Lu, H.; Cairney, J.; Chen, Y.-S. Hydrogen in pipeline steels: Recent advances in characterization and embrittlement mitigation. J. Nat. Gas Sci. Eng. 2022, 105, 104709. [Google Scholar] [CrossRef]
  30. Eames, I.; Austin, M.; Wojcik, A. Injection of gaseous hydrogen into a natural gas pipeline. Int. J. Hydrogen Energy 2022, 47, 25745–25754. [Google Scholar] [CrossRef]
  31. Abd, A.A.; Naji, S.Z.; Thian, T.C.; Othman, M.R. Evaluation of hydrogen concentration effect on the natural gas properties and flow performance. Int. J. Hydrogen Energy 2021, 46, 974–983. [Google Scholar] [CrossRef]
  32. Gönczi, G. Unconventional methods for pressure loss reduction in standard pipe elements. Water Pract. Technol. 2018, 13, 355–361. [Google Scholar] [CrossRef]
  33. Di Lullo, G.; Oni, A.O.; Kumar, A. Blending blue hydrogen with natural gas for direct consumption: Examining the effect of hydrogen concentration on transportation and well-to-combustion greenhouse gas emissions. Int. J. Hydrogen Energy 2021, 46, 19202–19216. [Google Scholar] [CrossRef]
  34. Vaccariello, E.; Trinchero, R.; Stievano, I.S.; Leone, P. A Statistical Assessment of Blending Hydrogen into Gas Networks. Energies 2021, 14, 5055. [Google Scholar] [CrossRef]
  35. Makaryan, I.A.; Sedov, I.V.; Salgansky, E.A.; Arutyunov, A.V.; Arutyunov, V.S. A Comprehensive Review on the Prospects of Using Hydrogen–Methane Blends: Challenges and Opportunities. Energies 2022, 15, 2265. [Google Scholar] [CrossRef]
  36. Shaaban, S. Design and optimization of a novel flowmeter for liquid hydrogen. Int. J. Hydrogen Energy 2017, 42, 14621–14632. [Google Scholar] [CrossRef]
  37. Njoka, F.; Ookawara, S.; Ahmed, M. Influence of design and operating conditions on the performance of tandem photoelectrochemical reactors. Int. J. Hydrogen Energy 2018, 43, 1285–1302. [Google Scholar] [CrossRef]
  38. Pozzi, A.; Tognaccini, R. The effect of the Eckert number on impulsively started pipe flow. Eur. J. Mech. B Fluids 2012, 36, 120–127. [Google Scholar] [CrossRef]
  39. Urbanowicz, K. Fast and accurate modelling of frictional transient pipe flow. ZAMM Z. Für Angew. Math. Mech. J. Appl. Math. Mech. 2018, 98, 802–823. [Google Scholar] [CrossRef]
  40. Dyachenko, S.A.; Zlotnik, A.; Korotkevich, A.O.; Chertkov, M. Operator splitting method for simulation of dynamic flows in natural gas pipeline networks. Phys. D. Nonlinear Phenom. 2017, 361, 1–11. [Google Scholar] [CrossRef]
  41. Bulckaen, V. Energy losses in the transport in 4000 km pipelines of liquid hydrogen and oxygen derived from the splitting of water, and of liquid methane. Int. J. Hydrogen Energy 1992, 17, 613–622. [Google Scholar] [CrossRef]
  42. Kürten, D.; Khader, I.; Kailer, A. Determining the effective hydrogen diffusion coefficient in 100Cr6. Mater. Corros. 2020, 71, 918–923. [Google Scholar] [CrossRef]
  43. Johnson, D.; Covington, A.; Clark, N. Environmental and Economic Assessment of Leak and Loss Audits at Natural Gas Compressor and Storage Facilities. Energy Technol. 2014, 2, 1027–1032. [Google Scholar] [CrossRef]
  44. Peet, Y.; Sagaut, P.; Charron, Y. Pressure loss reduction in hydrogen pipelines by surface restructuring. Int. J. Hydrogen Energy 2009, 34, 8964–8973. [Google Scholar] [CrossRef]
  45. Li, J.; Han, G.; Zhao, M.; Qu, W.; Nie, M.; Song, W.; Xie, B.; Eller, F. Nitrogen input weakens the control of inundation frequency on soil organic carbon loss in a tidal salt marsh. Estuar. Coast. Shelf Sci. 2020, 243, 106878. [Google Scholar] [CrossRef]
  46. Mohanty, S.; Nayak, A.K.; Swain, C.K.; Dhal, B.R.; Kumar, A.; Kumar, U.; Tripathi, R.; Shahid, M.; Behera, K.K. Impact of integrated nutrient management options on GHG emission, N loss and N use efficiency of low land rice. Soil Tillage Res. 2020, 200, 104616. [Google Scholar] [CrossRef]
  47. Zhuang, C.; Shao, J.; Wang, Z.; Lu, Y.; Zhang, K.; Dou, Z. Explosion suppression of porous materials in a pipe-connected spherical vessel. J. Loss Prev. Process Ind. 2020, 65, 104106. [Google Scholar] [CrossRef]
  48. Yuan, Z.; Fan, A. The effects of aspect ratio on CH4/air flame stability in rectangular mesoscale combustors. J. Energy Inst. 2020, 93, 792–801. [Google Scholar] [CrossRef]
  49. RoyChowdhury, T.; Bramer, L.; Hoyt, D.W.; Kim, Y.M.; Metz, T.O.; McCue, L.A.; Diefenderfer, H.L.; Jansson, J.K.; Bailey, V. Temporal dynamics of CO2 and CH4 loss potentials in response to rapid hydrological shifts in tidal freshwater wetland soils. Ecol. Eng. 2018, 114, 104–114. [Google Scholar] [CrossRef]
  50. Liu, Y.; Li, X.; Wang, W.; Li, L.; Huo, Y. Numerical investigation on the evolution of forces and energy features in thermo-sensitive cavitating flow. Eur. J. Mech. B Fluids 2020, 84, 233–249. [Google Scholar] [CrossRef]
  51. Sánchez-Orgaz, E.M.; Denia, F.D.; Baeza, L.; Kirby, R. Numerical mode matching for sound propagation in silencers with granular material. J. Comput. Appl. Math. 2019, 350, 233–246. [Google Scholar] [CrossRef]
  52. Dhamala, T.N.; Pyakurel, U.; Dempe, S. A Critical Survey on the Network Optimization Algorithms for Evacuation Planning Problems; TU Bergakademie Freiberg, Fakultät für Mathematik und Informatik: Leipzig, Germany, 2018; Volume 15, p. 101. [Google Scholar]
  53. Yang, L.; Ji, Z.L.; Wu, T.W. Transmission loss prediction of silencers by using combined boundary element method and point collocation approach. Eng. Anal. Bound. Elem. 2015, 61, 265–273. [Google Scholar] [CrossRef]
  54. Hemme, C.; van Berk, W. Hydrogeochemical modeling to identify potential risks of underground hydrogen storage in depleted gas fields. Appl. Sci. 2018, 8, 2282. [Google Scholar] [CrossRef]
  55. Lei, Y.; Hosseini, E.; Liu, L.; Scholes, C.A.; Kentish, S.E. Internal polymeric coating materials for preventing pipeline hydrogen embrittlement and a theoretical model of hydrogen diffusion through coated steel. Int. J. Hydrogen Energy 2022, 47, 31409–31419. [Google Scholar] [CrossRef]
  56. Ekhtiari, A.; Flynn, D.; Syron, E. Investigation of the Multi-Point Injection of Green Hydrogen from Curtailed Renewable Power into a Gas Network. Energies 2020, 13, 6047. [Google Scholar] [CrossRef]
  57. Eileen, S.; Melissa, K.V.G. 1 Existing Natural Gas Pipeline Materials and Associated Operational Characteristics. DOE Hydrog. Program 2006. Available online: https://www.hydrogen.energy.gov/docs/hydrogenprogramlibraries/pdfs/progress05/v_g_1_schmura.pdf?sfvrsn=79a23ec1_1.pdf (accessed on 12 August 2024).
  58. Huising, O.J.C.; Krom, A.H.M. H2 in an Existing Natural Gas Pipeline. In Volume 1: Pipeline and Facilities Integrity, 2020 13th International Pipeline Conference, Virtual, 28–30 September 2020; American Society of Mechanical Engineers: New York City, NY, USA, 2020. [Google Scholar] [CrossRef]
  59. Injecting Hydrogen in Natural Gas Grids Could Provide Steady Demand the Sector Needs to DeveloElectric Power. Natural Gas. SP Global, 19 May 2020. Available online: https://www.spglobal.com/platts/en/market-insights/blogs/natural-gas/051920-injecting-hydrogen-in-natural-gas-grids-could-provide-steady-demand-the-sector-needs-to-develop (accessed on 12 August 2024).
  60. Tong, S.; Li, X.; Sun, S.; Tu, C.; Xia, X. Interchangeability of Hydrogen Injection in Zhejiang Natural Gas Pipelines as a Means to Achieve Carbon Neutrality. Energies 2022, 15, 6394. [Google Scholar] [CrossRef]
  61. Tsui, L.; Garzon, F.; Agi, K. Solid-State Mixed-Potential Electrochemical Sensors for Natural Gas Leak Detection and Quality Control (Final Technical Report); University of New Mexico: Albuquerque, NM, USA, 2024; DOE-UNM--FE0031864; p. 2382681. [Google Scholar] [CrossRef]
  62. Mercuri, A.; Gianfelici, F.; Blasioli, G.; Luci, V.; Branduardi, L.; Arcangeletti, G.; Aloigi, E. Safe Delivery of H2 and CO2 in Offshore Pipeline Systems: Novel Methodology and Tools for Technological Risk Assessment. In Proceedings of the Offshore Technology Conference, OTC, Houston, TX, USA, 2–5 May 2022. [Google Scholar] [CrossRef]
  63. Isaac, T. HyDeploy: The UK’s First Hydrogen Blending Deployment Project. Clean Energy 2019, 3, 114–125. [Google Scholar] [CrossRef]
  64. Mark, F.R.; Paige, J.; Nicholas, G.; Elizabeth, C.; Richard, B.; Simon, A.J.; Amgad, E.; Jarett, Z. The Technical and Economic Potential of the H2@Scale Concept within the United States; National Renewable Energy Laborator: Golden, CO, USA, 2020; NREL/TP-6A20-77610. Available online: https://www.nrel.gov/docs/fy21osti/77610.pdf (accessed on 12 August 2024).
  65. Lamari, F.; Weinberger, B.; Langlois, P.; Fruchart, D. Instances of Safety-Related Advances in Hydrogen as Regards Its Gaseous Transport and Buffer Storage and Its Solid-State Storage. Hydrogen 2024, 5, 387–402. [Google Scholar] [CrossRef]
  66. Smith, J. The Role of Hydrogen Fuel in Hard-to-Decarbonise Modes of Transport: An Energy Systems Perspective. Ph.D. Thesis, Apollo—University of Cambridge Repository, Cambridge, UK, 2022. [Google Scholar] [CrossRef]
  67. Klopčič, N.; Stöhr, T.; Grimmer, I.; Sartory, M.; Trattner, A. Refurbishment of Natural Gas Pipelines towards 100% Hydrogen—A Thermodynamic-Based Analysis. Energies 2022, 15, 9370. [Google Scholar] [CrossRef]
  68. Zhang, J.X.; An, C.; Wei, D.F.; Chen, B.Q.; Soares, C.G. Structural Behaviour of Hydrogen Flexible Pipe under Internal Pressure. In Trends in Renewable Energies Offshore; CRC Press: Boca Raton, FL, USA, 2022. [Google Scholar]
  69. Victor FAGNON. Techno-Economic Assessment of Flexible Composite Pipelines for Decentralised Production of H2 with Electrolysers Integrated in Wind Turbine of Stavanger, 2022. Available online: https://uis.brage.unit.no/uis-xmlui/bitstream/handle/11250/3032550/no.uis:inspera:102983723:64696369.pdf?sequence=1 (accessed on 12 August 2024).
  70. Nemeth, A.; Czapski, D.; Alexander, C. Optimizing Operator Systems Through the Use of Flexible Composite Pipe. In Volume 2: Pipeline and Facilities Integrity; American Society of Mechanical Engineers: Calgary, AB, Canada, 2022. [Google Scholar] [CrossRef]
  71. Zhang, C.; Shao, Y.; Shen, W.; Li, H.; Nan, Z.; Dong, M.; Bian, J.; Cao, X. Key Technologies of Pure Hydrogen and Hydrogen-Mixed Natural Gas Pipeline Transportation. ACS Omega 2023, 8, 19212–19222. [Google Scholar] [CrossRef] [PubMed]
  72. Tan, K.; Mahajan, D.; Venkatesh, T.A. Computational fluid dynamic modeling of methane-hydrogen mixture transportation in pipelines: Understanding the effects of pipe roughness, pipe diameter and pipe bends. Int. J. Hydrogen Energy 2024, 49, 1028–1042. [Google Scholar] [CrossRef]
  73. Thawani, B.; Hazael, R.; Critchley, R. Assessing the pressure losses during hydrogen transport in the current natural gas infrastructure using numerical modelling. Int. J. Hydrogen Energy 2023, 48, 34463–34475. [Google Scholar] [CrossRef]
  74. Ruiz-Tagle, A.; Groth, K.M. Comparing the risk of third-party excavation damage between natural gas and hydrogen pipelines. Int. J. Hydrogen Energy 2024, 57, 107–120. [Google Scholar] [CrossRef]
Figure 1. Processes of hydrogen generation efficiency and costs [10].
Figure 1. Processes of hydrogen generation efficiency and costs [10].
Energies 17 04709 g001
Figure 2. Flow chart of the systematic review process flow.
Figure 2. Flow chart of the systematic review process flow.
Energies 17 04709 g002aEnergies 17 04709 g002b
Figure 3. Identified working pressure of the transported hythane (hydrogen–methane).
Figure 3. Identified working pressure of the transported hythane (hydrogen–methane).
Energies 17 04709 g003
Figure 4. Parameters of pipeline networks for hythane transport.
Figure 4. Parameters of pipeline networks for hythane transport.
Energies 17 04709 g004
Figure 5. Material used in pipeline networks for hythane transport.
Figure 5. Material used in pipeline networks for hythane transport.
Energies 17 04709 g005
Figure 6. Parameters of hythane transport in network pipes included in review.
Figure 6. Parameters of hythane transport in network pipes included in review.
Energies 17 04709 g006
Figure 7. Maximum acceptable H2 percentage in CH4 transport systems across different countries.
Figure 7. Maximum acceptable H2 percentage in CH4 transport systems across different countries.
Energies 17 04709 g007
Figure 8. Timeline of accidents related to CH4 (upper chart) and H2 (lower chart) based on fatalities, damages, and media coverage.
Figure 8. Timeline of accidents related to CH4 (upper chart) and H2 (lower chart) based on fatalities, damages, and media coverage.
Energies 17 04709 g008
Table 1. Search criteria used to populate the review research papers’ database.
Table 1. Search criteria used to populate the review research papers’ database.
Criterion 1Criterion 2Criterion 3Criterion 4Criterion 5Criterion 6
Key wordsH2 losses in pipesLosses in pipesFluid transmission losses Hythane losses CH4 lossesNA 1
DisciplinesAppendix AAll
ExpandersAlso search within the full text of the articles/apply equivalent subjectsNA
Search modesAppendix AAppendix A
Results’ limitsFull text peer-reviewed
Published date2012 ÷ 2024No time limit
LanguageEnglish
Result [papers]132004034885
1 The sixth criterion refers to the papers that were recommended by online page recommendation during the full text review of the research papers defined in criteria 1 to 5.
Table 2. Physical proprieties of H2, CH4, and mixtures (volume %).
Table 2. Physical proprieties of H2, CH4, and mixtures (volume %).
PropertiesHydrogen (H2)Methane (CH4)Hythane
10% H2-90% CH4
Hythane
25% H2-75% CH4
Hythane
30% H2-70% CH4
Mixture 40% H2-60% CH4Mixture 60% H2-40% CH4Unit
Lower calorific value by volume at 1 atm11.0035.0032.6027.2033.3225.4020.60MJ/m3
Higher calorific value by volume at 1 atm13.0039.0036.0030.0027.0028.0020.00MJ/m3
Maximum flame temperature1800.001495.001525.501594.131516.351617.001678.00K
Auto-ignition temperature560.00600.00590.00590.00580.00580.00570.00C
Laminar burning velocity3.100.400.681.310.601.522.08m/s
Table 3. Risk matrix for H2 utilization in existing methane pipelines.
Table 3. Risk matrix for H2 utilization in existing methane pipelines.
ConsequencesOccurrence Probability
0–20%21–40%41–60%61–80%81–100%
ABCDE
SeverityHealth/SafetyEnvironmentPRCorporate
Impact
Almost CertainRareUnlikelyPossibleLikelyCertain
5
Severe
Potential for multiple fatalitiesPotential catastrophic damageGovernmental levelLack of functionality >7 days>3 days5A5B5C5D5E
4
Major
Potential for single fatalityPotential long-term effectsPublic disruptionLack of functionality >1 day<3 days4A4B4C4D4E
3
Moderate
Potential for a single major injuryPotential medium-term effectsSmall public disruptionLack of functionality <1 day>2 days3A3B3C3D3E
2
Minor
Potential for lost timePotential long-term effectsLocal media coverageCorrective maintenance>2 days2A2B2C2D2E
1
Marginal
Potential for first aid injuryPotential long-term effectsNo media coverageMalfunctionNo imp.1A1B1C1D1E
Table 4. Maximum allowable H2 mixture in CH4 transport networks.
Table 4. Maximum allowable H2 mixture in CH4 transport networks.
CH4 InfrastructurePressure RegulationMetersCHG Storage TanksHouse InstallSeals
/Valves
Transmission PipelinesCo-Gen PlantsHome Gas Burners/StoveCompression StationGas Turbines
H2 [%]673030303030201021
Transport networkSteel pipesPVC pipelineSealantsConnectorsFittingsFlow valvesDomestic pipe
H2 [%]30703030301530
Transport equipmentPipelineTurbineCompressor
3015
Table 5. Comparative properties of methane and hydrogen relevant to transport and safety.
Table 5. Comparative properties of methane and hydrogen relevant to transport and safety.
MethaneHydrogen
greater densityhigher heat capacity
greater viscosityhigher diffusivity
higher calorific value by volumehigher calorific value by mass
higher solubility in waterhigher flame temperature
higher auto-ignition temperature
wider explosive and fire danger
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hora, C.; Dan, F.C.; Secui, D.-C.; Hora, H.N. Systematic Literature Review on Pipeline Transport Losses of Hydrogen, Methane, and Their Mixture, Hythane. Energies 2024, 17, 4709. https://doi.org/10.3390/en17184709

AMA Style

Hora C, Dan FC, Secui D-C, Hora HN. Systematic Literature Review on Pipeline Transport Losses of Hydrogen, Methane, and Their Mixture, Hythane. Energies. 2024; 17(18):4709. https://doi.org/10.3390/en17184709

Chicago/Turabian Style

Hora, Cristina, Florin Ciprian Dan, Dinu-Calin Secui, and Horea Nicolae Hora. 2024. "Systematic Literature Review on Pipeline Transport Losses of Hydrogen, Methane, and Their Mixture, Hythane" Energies 17, no. 18: 4709. https://doi.org/10.3390/en17184709

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop