Next Article in Journal
Estimation of Remaining Insulation Lifetime of Aged XLPE Cables with Step-Stress Method Based on Physical-Driven Model
Previous Article in Journal
Study on the Characteristics of High-Temperature and High-Pressure Spray Flash Evaporation for Zero-Liquid Discharge of Desulfurization Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Simulation Study of Gas Stratification in Hydrogen-Enriched Natural Gas Pipelines

1
State Key Laboratory of Oil and Gas Reservoir Geology and Engineering, Southwest Petroleum University, Chengdu 610500, China
2
School of New Energy and Materials, Southwest Petroleum University, Chengdu 610500, China
3
Southwest Branch Company, China Petroleum Engineering Construction Corporation, Chengdu 610031, China
*
Author to whom correspondence should be addressed.
Energies 2025, 18(12), 3181; https://doi.org/10.3390/en18123181
Submission received: 17 May 2025 / Revised: 10 June 2025 / Accepted: 12 June 2025 / Published: 17 June 2025

Abstract

:
Hydrogen blending in natural gas pipelines facilitates renewable energy integration and cost-effective hydrogen transport. Due to hydrogen’s lower density and higher leakage potential compared to natural gas, understanding hydrogen concentration distribution is critical. This study employs ANSYS Fluent 2022 R1 with a realizable k-ε model to analyze flow dynamics of hydrogen–methane mixtures in horizontal and undulating pipelines. The effects of hydrogen blending ratios, pressure (3–8 MPa), and pipeline geometry were systematically investigated. Results indicate that in horizontal pipelines, hydrogen concentrations stabilize near initial values across pressure variations, with minimal deviation (maximum increase: 1.6%). In undulating pipelines, increased span length of elevated sections reduces maximum hydrogen concentration while maintaining proximity (maximum increase: 0.65%) to initial levels under constant pressure. Monitoring points exhibit concentration fluctuations with changing pipeline parameters, though no persistent stratification occurs. However, increasing the undulating height elevation difference leads to an increase in the maximum hydrogen concentration at the top of the pipeline, rising from 3.74% to 9.98%. The findings provide theoretical insights for safety assessments of hydrogen–natural gas co-transport and practical guidance for pipeline design optimization.

1. Introduction

The global intensification of energy crises coupled with escalating environmental deterioration has rendered the establishment of sustainable energy systems implementing carbon-neutral technologies an imperative societal objective [1]. Hydrogen emerges as a leading next-generation energy vector, distinguished by its combustion producing water vapor without greenhouse gas emissions or particulate matter generation. With a gravimetric energy density of 143 MJ·kg−1 [2,3], this lightweight gas delivers specific energy content exceeding conventional liquid hydrocarbons (diesel/petrol) by nearly threefold. Critical technological impediments persist, however, in establishing economically viable storage and transportation infrastructure that ensures both operational safety and material integrity.
The escalating hydrogen economy demands have driven comprehensive multidisciplinary investigations into retrofitting fossil-derived gas transportation frameworks with hydrogen compatibility protocols [4,5,6]. This cross-domain synergy leverages legacy pipeline assets’ structural sunk capital while maintaining hydraulic integrity, simultaneously achieving two essential energy transition objectives: infrastructure interoperability preservation and hydrogen vectorization at gigawatt-hour scalability thresholds [7,8].
The rapid expansion of China’s natural gas pipeline infrastructure has heightened scientific attention to operational shutdown events [9,10], known to induce significant thermodynamic variations including temperature and pressure fluctuations within pipeline systems [11,12,13]. While extensive research has addressed critical operational challenges such as blockage mitigation, shutdown duration safety thresholds, and thermal dissipation patterns, scientific investigation of hydrogen-enriched natural gas pipeline shutdown phenomena remains comparatively underdeveloped. Current experimental studies primarily focus on density-driven stratification dynamics, demonstrating that gas mixtures with substantial density differentials [14] inherently develop compositional stratification—a process independent of pressure conditions [15], thermal gradients [16], or external environmental factors [17,18,19]. This stratification mechanism, exacerbated by extended transport distances and shutdown durations, raises material safety concerns through localized hydrogen accumulation in pipeline systems [20,21,22].
Numerous researchers have investigated gas mixtures under gravitational influence; however, considerable controversy still exists regarding the stratification behavior of hydrogen–methane mixtures. Tong [23] established a 2.96 m (width) × 2.24 m (height) one-dimensional model to investigate gas distribution under varying hydrogen fluxes. Results indicate more uniform hydrogen distribution at low Froude numbers where stratification is gravity-dominated. Peng et al. [24] systematically studied equilibrium distribution of hydrogen–methane mixtures in gravitational fields by integrating molecular dynamics simulations with thermodynamics and diffusion theory. Their findings demonstrate that partial pressures of hydrogen and methane decrease with increasing height. Due to its lower molar mass, hydrogen exhibits a slower decrease rate, leading to accumulation at the pipeline top. However, significant stratification occurs only under substantial height differences, with stabilization time increasing proportionally to the height differential. Pitts et al. [25] constructed a two-car garage chamber to explore hydrogen dispersion following vehicular leaks. After ground-level central release, a hydrogen concentration gradient forms with peak concentrations near the ceiling, though concentration differences remain minor. Collectively, stratification of hydrogen-methane mixtures in gravitational fields remains contentious, with limited studies in horizontal/undulating natural gas pipelines, necessitating systematic quantitative investigations of stratification behavior.
Under conventional operating conditions, the gas flow in pipelines is predominantly turbulent, which promotes the thorough mixing of natural gas and hydrogen, preventing stratification. However, once the pipeline is shut down for maintenance or repair, the situation changes. Due to the significant density differences between hydrogen and natural gas, hydrogen tends to separate from the mixture and accumulate at the top when the gas is at rest, leading to stratification. If this stratification results in a sharp increase in hydrogen concentration at the top, it significantly raises the risk of hydrogen-induced cracking in the pipeline, posing a safety threat.
Therefore, understanding the stratification mechanism, the spatiotemporal variation of hydrogen concentration, and the time required to reach a stable stratified state during the gas standing process after pipeline shutdown is critical for determining the safety maintenance window of hydrogen-blended pipelines. Unfortunately, current research in this area is insufficient both domestically and internationally, and further exploration is urgently needed to fill this knowledge gap.

2. Research Background and Current Status of the Study

When considering the integration of hydrogen into natural gas pipeline systems for transportation, the primary task is to determine an appropriate hydrogen blending ratio. However, this is not a straightforward process, as it is influenced by numerous complex factors, including but not limited to pipeline material compatibility, changes in the physical properties of the gas mixture, and the adaptability of end-use equipment. As a result, no global unified standard has been established to date. Different countries, based on their own technical conditions, safety regulations, and economic considerations, have set varying thresholds for the upper limits of hydrogen blending. For instance, in Europe [26], Finland has taken a conservative approach, setting the upper limit at 1%, while Switzerland [27], Austria [28], and Spain [29,30] have raised this value to 2%, 4%, and 5%, respectively, showing a more open attitude. Meanwhile, a study by the Australian Renewable Energy Agency indicates that when the hydrogen blending ratio is kept below 10% [31], its impact on existing natural gas pipeline infrastructure, end-use equipment, and current regulatory systems is negligible or minimal, providing the industry with operational flexibility and a reference point.
In China, significant strides are being made in the utilization of hydrogen energy. For example, the “Chaoyang Renewable Energy Hydrogen Blending Demonstration Project Phase I” successfully achieved a hydrogen blending ratio of approximately 5%, accumulating valuable experience for hydrogen transportation technology in the country. However, it is important to note that China has not yet issued clear laws, regulations, or technical standards regarding the upper limits of hydrogen blending in natural gas pipelines and networks. This presents both a wide space for industry innovation and a need for continuous learning from practice, gradually establishing a comprehensive standard system to guide the healthy and orderly development of the hydrogen energy industry [32].

2.1. Hydrogen-Blended Natural Gas Pipeline Transportation Cases

Although the development of hydrogen pipelines in China started relatively late, it has been accelerating in recent years. The total length of the hydrogen pipeline network currently approaches 400 km, primarily concentrated in the Bohai Economic Rim and Yangtze River Delta regions, reflecting a regionalized layout. Notably, the hydrogen pipeline connecting Jiyuan City and Luoyang City in Henan Province marks a significant breakthrough in China’s hydrogen transportation technology. With a length of 25 km, an oversized diameter of 508 mm, a hydrogen transport pressure of up to 4 MPa, and an annual hydrogen transport capacity exceeding 100, 400 tons, this pipeline has become a leader in the field of hydrogen pipeline construction in China.
Internationally, the development of hydrogen pipelines has a long history, dating back to the late 1930s, marking the early emergence of hydrogen pipeline transportation technology. Although the total length of hydrogen pipelines globally has accumulated to approximately 4500 km, this figure is still relatively small compared to the vast network of oil and gas pipelines, showing a significant gap. Focusing on Europe, by the end of 2016, about 1598 km of hydrogen pipelines had been laid, with operating pressures typically ranging from 2 to 10 MPa. These pipelines are mainly made of seamless steel, with diameters ranging from 0.3 to 1.0 m. In terms of material selection, pipeline steels such as X42, X52, and X56, which are low-strength but highly adaptable, have become mainstream. Notably, Germany has a long history of hydrogen pipeline construction. As early as 1939, Germany constructed a 208 km-long hydrogen pipeline with a 254 mm diameter and a 2 MPa operating pressure, achieving a hydrogen transport rate of 9000 kg per h. This achievement not only highlighted Germany’s pioneering position in hydrogen energy technology but also set a benchmark for the global development of hydrogen pipelines.
Table 1 provides detailed information on the specific lengths of hydrogen pipelines in different regions, based on statistics available as of the end of 2016. These figures offer a quantitative reference for the development trajectory of the hydrogen energy sector. The primary owners of major hydrogen pipelines are Liquefied Gas Companies and Air Products, with respective pipeline lengths of approximately 1936 km and 1140 km. Detailed data can be found in Table 2.

2.2. Theoretical Analysis of Distribution in Mixed Gas Systems at Rest

From the perspective of the second law of thermodynamics, the process of non-uniform concentration distribution of gases in a pipeline from a fully mixed state is a process of entropy reduction and does not occur spontaneously. However, in reality, the gases in the pipeline are not ideal gases and are affected by the gravitational field. Due to the differences in molecular weights of the components in the gas mixture, a concentration gradient can develop in the vertical direction. The existence of the thin upper atmosphere on Earth serves as proof of this possibility.
To study the macroscopic motion of a large number of particles, Boltzmann proposed the highly influential principle of equiprobability, which forms the foundation of classical statistical mechanics. This principle asserts that when an isolated system reaches equilibrium, all possible microscopic states occur with equal probability. Therefore, we can derive the most probable distribution at the microscopic level—the Maxwell–Boltzmann distribution.
For typical building heights near the Earth’s surface, it is assumed that the temperature remains constant at room temperature (25 °C or 298 K). Taking a mixture of 20% hydrogen and 80% methane as an example, the concentration ratio of the two gases along height variations is shown in Table 3.
As seen in the table, with the increase in height, the proportion of the lighter H2 in the mixed gas increases, while the volumetric fraction of the heavier gas, CH4, decreases. Within typical building heights, the change in the concentration ratio of the mixed gas is less than 1%. Specifically, even for a mixed gas column at a height of 1 km, the change in the concentration ratio of the components is only about 1%. Scholars who have studied these issues have all adopted the aforementioned derivations. V.V. Azatyan et al. [33] studied the gravitational stratification possibility of a propane-hydrogen mixture, concluding that the gravity distribution of the gases follows the Boltzmann distribution. Based on this, they derived the variation in the concentration ratio of hydrogen and propane with height, noting that under static conditions at a height of 15 m, the change in the concentration ratio was less than 0.2%. If natural convection is considered, the difference becomes even smaller.
It is widely believed that gas stratification occurs in areas with little flow, such as caves and sewers filled with CO2 in nature. Badino G [34] provided mathematical analysis and theoretical derivation of the behavior of such gases in confined conditions, suggesting that the stratification of underground gas layers with specific components is not due to gravity, but rather because there is no air movement around the gas source. Stratification does occur, but it requires a static air column several kilometers high to have an effect. The occurrence of stratification in nature is not due to density differences, but because these gases are constantly generated in caves and diffuse very slowly, leading to localized concentration at certain heights. In high-altitude non-homogeneous layers, where the temperature remains constant, there is a tendency for gas stratification. In higher layers, components with smaller molar masses have higher concentrations, and when the height is sufficiently high, the components with larger molar masses may even disappear, though this occurs at altitudes of tens of kilometers.

3. Numerical Simulation Methodology

Due to the complexity of the gas mixing process, it is difficult to measure the concentration changes at various points. CFD (Computational Fluid Dynamics) simulation has become a promising approach to address this issue. Some scholars argue that due to the influence of gravity, when gases with different densities are mixed, concentration non-uniformity occurs. Other scholars, however, adhering to the theory of molecular thermal motion, believe that stratification is unlikely to occur on a small scale. These issues can be explored, observed, and studied using advanced and mature CFD tools, while small-scale experiments can be conducted for validation. However, it should be noted that since CFD software does not operate at the molecular scale, the simulation results tend to be more conservative.

3.1. Governing Equations

Considering gas compressibility, the unsteady Reynolds-Averaged Navier–Stokes (RANS) equations are employed to solve the mixing process of hydrogen and natural gas in high-pressure pipelines [35]. The Boussinesq eddy-viscosity approach establishes the relationship between Reynolds stresses and mean velocity gradients [36]:
    τ i j = ρ v m ,   i ,   v m ,   j   ,   = 2 μ t S i j 2 3 ρ k δ i j    
where
μ t = C μ ρ k 2 ε , S i j = 1 2 v m ,   i x j + v m ,   j x i
where τij is Reynolds stress, Pa, i and j is coordinate component indices (I, j = 1, 2, 3), k and ε represent turbulent kinetic energy and turbulent energy dissipation rate, respectively, Sij is strain, μt is the eddy viscosity coefficient, Cμ is the empirical constant, Cμ = 0.09, δij is kronecker delta, vm is mixture gas velocity, m/s, ρ is the gas density, kg/m3.
Studies indicate that the Reynolds Stress model is suitable for three-dimensional complex flows but exhibits poor convergence and requires greater computational time and memory. The Standard k-ε turbulence model yields distorted results when applied to curved surface flows. Both Realizable and RNG k-ε models address this limitation; however, compared to RNG, the Realizable k-ε turbulence model demonstrates broader applicability in non-separated flow problems with superior convergence and lower computational costs [37,38]. The realizable k-ε turbulence model closes the governing equations, including:
ρ m k t + ρ m v m ,   j k x j = τ i j v m ,   j x j + x j μ m +   μ t k x j ρ m ε      
ρ m ε t + ρ m v m ,   j ε x j = C ε 1 τ i j ε v m ,   j k x j + x j μ m + μ t σ ε ε x j + C ε 2 ρ m ε 2 k  
where t is time, s, ρm is the density of the mixed gas, kg/m3, μm is the dynamic viscosity of the mixed gas, Pa∙s, σε, Cε1, Cε2 are empirical constants, taken as σε = 1.3, Cε1 = 1.45, Cε2 = 1.90.
For multiphase flow modeling: The VOF model requires distinct interfaces between fluids, rendering it unsuitable for hydrogen–natural gas mixing simulations. The Euler model entails excessive computational demands. Consequently, the Mixture model—applicable to homogeneous dispersed multiphase flows—is selected. The governing equations for hydrogen–natural gas stratification solved by the Mixture model are as follows [35]:
      ρ m t + ρ m v m = 0
ρ m v m t + · ρ m v m × v m = p + · μ m v m + v m T                                                                                                                                                                                                 + ρ m g + · k = 1 N α k ρ k v d r , k × v d r , k                                                                                                                    
where
ρ m = k = 1 N α k ρ k , v m = k = 1 N α k ρ k v d r , k / ρ m , v d r , k = v k v m ,
= i x + j x + k x , μ m = k = 1 N α k μ k
where ρk is the k-th phase density, kg/m3; Vm and Vk are the average velocity of the mixture and the velocity of the k-th phase, respectively, m/s; αk is the volume fraction of the k-th phase (H2 and natural gas); N is the number of phases; μk is the k-th phase dynamic viscosity, Pa∙s; dr, Vk is the drift velocity, m/s; P is static pressure, Pa; g is gravitational acceleration, m/s2.

3.2. Problem Description and Grid Generation

The computational domain used in the simulation is a horizontal pipeline with a length of 30 m and a diameter of 500 mm. Three monitoring points are set along the pipe axis: P1 (0, 0245, 15), P2 (0, 0, 15), and P3 (0, −0245, 15) to monitor the hydrogen concentration distribution along the pipeline. The geometric diagram is shown in Figure 1.
During the long-distance transportation of natural gas through pipelines, complex terrain conditions, such as mountains and rivers, may be encountered. This chapter focuses on the study of static stratification in undulating natural gas pipelines that may be encountered in actual field conditions. The two-dimensional model of the undulating pipeline used in this study is shown in Figure 2.
Static stratification simulations in horizontal pipelines investigate the effects of internal pressure and initial hydrogen concentration (volume fraction, φH2,I) on hydrogen-natural gas mixtures, determining whether stratification occurs and quantifying its patterns. Simulations in undulating pipelines conduct sensitivity analyses of hydrogen concentration, elevation difference (H), top span length (L2), and number of spans (n) to establish stratification patterns in undulating pipelines. Detailed simulation conditions are listed in Table 4.
Structured meshing was applied to the geometric models shown in Figure 1 and Figure 2, as illustrated in Figure 3 and Figure 4, with local refinement implemented at bends of the undulating pipeline. All boundaries were set as stationary walls with the no-slip condition. A grid independence study was conducted to eliminate the influence of mesh density on simulation results. For an internal pressure of 8 MPa and 3% hydrogen concentration, stabilized hydrogen concentrations for models with different mesh counts are presented in Table 5. The data demonstrate that for static stratification in horizontal pipelines, hydrogen concentration stabilizes when the mesh count exceeds 250,000. Balancing computational accuracy and efficiency, the mesh count was set to 254,566. For static stratification in undulating pipelines, hydrogen concentration remains nearly constant above 400,000 elements. Considering computational precision and efficiency, the mesh count was set to 405,612.

3.3. Model Validation

At present, there are relatively few experimental studies on the stratification of H2–methane mixed gas. Ren [39] injected methane into a sealed tank filled with air and observed vertical stratification of the gas mixture after static placement. Ren employed a sealed vessel (height: 0.33 m; diameter: 0.14 m). Initial internal pressure was 0.1 MPa. Following the partial pressure principle, a gas mixture was introduced to establish initial methane volume fractions of 5%, 10%, and 15%. Methane volume fractions at different locations were measured after an equilibration period. The numerical model was validated against these experimental data using identical operational parameters. As shown in Figure 5, simulated results demonstrate good agreement with experimental measurements. The simulated concentrations were overestimated as only gravitational effects were considered without accounting for thermal molecular motion, but a maximum deviation of 4.75% was observed between simulated and experimental values. Computational results demonstrate good agreement with experimental data. The numerical simulation methodology employed in this study accurately captures the dynamic process of gas stratification. This confirms the model’s capability to effectively simulate mixing-diffusion and stratification processes in H2-blending natural gas systems.

4. Study on the Static Concentration Distribution Pattern of Medium in Horizontal Pipelines

4.1. Static Stratification in a Horizontal Pipe at 3 MPa Internal Pressure

Based on the above numerical model, a simulation analysis of stratification in a hydrogen-blended pipeline under static conditions was conducted. The simulation conditions included internal pipe pressures of 3 MPa, with an internal pipe temperature of 300 K. The stratification phenomenon in horizontal pipes was compared for different hydrogen concentrations in natural gas, specifically 3%, 5%, 10%, and 50%.
Figure 6 shows the variation of hydrogen concentration along the vertical line at different time intervals (t = 400 s, 800 s, 1200 s, 1600 s, 2000 s, 4000 s, 6000 s, 10,000 s, 2000 s) in horizontal pipes with different initial hydrogen concentrations. It is evident that as time progresses, the hydrogen concentration at the top of the pipeline increases, while the concentration at the bottom decreases. At t = 2000 s, the rate of increase in hydrogen concentration accelerates, and around t = 10,000 s, the hydrogen concentration reaches a stable value. Figure 7 presents the variation of hydrogen concentration over time at the z = 15 m cross-section in horizontal pipes with different initial hydrogen concentrations.
Table 6 displays the static stratification results for the horizontal pipe at 3 MPa internal pressure. As the hydrogen blending ratio increases, the percentage increase in hydrogen concentration at the top of the pipe after stabilization becomes smaller. The final stabilization time shows no significant change, fluctuating around t = 10,000 s, and exhibits minimal influence from the initial concentration.

4.2. Static Stratification in Horizontal Pipes at 5 MPa and 8 MPa Internal Pressures

As the variation of hydrogen concentration along the vertical line and the variation pattern of hydrogen concentration over time at the z = 15 m cross-section in horizontal pipes exhibit consistent patterns across different pressures and hydrogen concentrations, additional figures are not provided. Figure 6 and Figure 7 can be referred to.
Table 7 presents the static stratification results for horizontal pipes at 5 MPa and 8 MPa internal pressures. With increasing pressure, the final stabilized concentration shows a slight increase. That is because increased pressure enlarges the density difference between hydrogen and methane, intensifying the gravity-driven segregation phenomenon. Additionally, hydrogen’s partial molar volume advantage becomes more pronounced under high pressure, promoting its migration towards the top.
The final stabilization time exhibits no significant change, occurring at approximately t = 10,000 s, and is minimally affected by the initial concentration and internal pressure. Although increased pressure reduces the diffusion coefficient, potentially prolonging the time to reach equilibrium, the enhanced natural convection due to the larger density difference under high pressure accelerates the upward movement of hydrogen and downward movement of methane, promoting segregation and shortening the equilibrium time. These two effects counteract each other, resulting in similar equilibrium times.

5. Study on the Static Concentration Distribution of Medium in Undulating Pipelines

5.1. Effect of Elevation Difference on Static Stratification in Undulating Pipelines

Based on the aforementioned numerical model, a simulation analysis of the stratification in undulating pipes was conducted. The simulation conditions are as follows:
Analysis was conducted using an elevation difference of 10 m. Figure 8 shows the hydrogen concentration distribution on the vertical cross-section of the undulating pipeline at nine time points: t = 1000 s, 2000 s, 4000 s, 10,000 s, 30,000 s, 50,000 s, 100,000 s, 150,000 s, and 280,000 s. The contour plot of hydrogen concentration along the longitudinal section reveals that hydrogen initially accumulates at the top of the undulating pipeline while decreasing gradually at the bottom over standing time. This occurs because gas molecules migrate downward under gravity, with molecular density decreasing exponentially with height, resulting in a non-uniform spatial distribution characterized by a sparse upper and dense lower profile. After 280,000 s (3.24 days), the hydrogen concentration at the pipeline top reaches 3.74%, representing a 24.6% increase from the initial state.
Figure 9 displays the hydrogen concentration at cross-sections of the upper and lower horizontal segments in the 10 m elevation difference pipeline. Within the first 10,000 s of standing, no stratification occurs between hydrogen and natural gas, with both gases distributed uniformly throughout the pipeline. As standing time increases further, the hydrogen concentration at the top gradually rises. Figure 10 presents concentration curves at monitoring points within the undulating pipeline. With prolonged standing time, concentration increases at the top monitoring point while decreasing at the bottom point. After 280,000 s (3.24 days), the maximum hydrogen concentration reaches 3.74% at the monitoring points.
Table 8 shows simulation results for different elevation differences under a fixed top span length (L2 = 10 m). The data indicate that when operational conditions and pipeline geometry remain unchanged, increasing the elevation difference leads to higher maximum hydrogen concentrations at the pipeline top after 3 days of standing—rising from 3.74% at 10 m to 9.98% at 50 m. This is because within the inclined sections of an undulating pipeline, shear-induced airflow moving upwards and downwards is formed due to the velocity difference between gas near the wall and at the center. An increase in the elevation difference intensifies this gas shearing effect, leading to more pronounced hydrogen concentration accumulation across the crown and thereby increasing the hydrogen concentration.

5.2. Effect of Top Span Length on Static Stratification in Undulating Pipelines

Analysis was conducted using a top span length of 50 m (Figure 11). Figure 12 shows the hydrogen concentration distribution on the vertical cross-section of the undulating pipeline at ten time points: t = 1000 s, 2000 s, 4000 s, 10,000 s, 30,000 s, 50,000 s, 100,000 s, 150,000 s, 200,000 s, and 280,000 s (top span length: 50 m). With increasing standing time, hydrogen concentration gradually rises at the pipeline top while decreasing at the bottom. Due to the density difference between methane and hydrogen molecules, molecular density decreases exponentially with height withFin the undulating pipeline, resulting in a non-uniform spatial distribution characterized by a sparse upper and dense lower profile. At t= 280,000 s (3.24 days), the hydrogen concentration at the pipeline top reaches 10.45%, representing a 0.45% increase from the initial state.
Figure 13 displays hydrogen concentration curves at monitoring points on cross-sections of the upper (points b, c) and lower (points a, d) horizontal segments (top span length: 50 m). Concentration increases over time at upper monitoring points but decreases at lower points, with the rate of concentration increase diminishing over time. The trend of hydrogen concentration change at point a is consistent with that at point d. After 280,000 s (3.24 days), the maximum hydrogen concentration (10.45%) occurs at upper monitoring points (Figure 14).
Under the condition of constant internal pressure and initial hydrogen concentration, as the top span length (L2) increases, the maximum hydrogen concentration at the top of the undulating pipe tends to decrease after 3 days of settling, from 10.65% at a span length of 10 m to 10.38% at 100 m (Table 9). A larger top span length in an undulating pipeline allows for more complete flow development, leading to reduced hydrogen concentration at the crown. Conversely, with a shorter top span length, disturbances induced by the inclined pipe sections do not fully dissipate before entering the opposing inclined section, resulting in a short residence time for disturbances and a more rapid initiation of gravity-driven stratification. In configurations with a larger top span length, disturbances generated by the inclined sections persist longer within the top pipe section or may even dissipate there, inducing more vigorous fluid disturbances. This increases the degree of gas mixing and consequently lowers the hydrogen concentration at the crown.

5.3. Effect of Number of Spans on Static Stratification in Undulating Pipelines

An analysis was conducted for n = 3 spans (Figure 15). Figure 16 shows hydrogen concentration on the vertical cross-section at ten time points: t = 1000 s, 2000 s, 4000 s, 10,000 s, 30,000 s, 50,000 s, 100,000 s, 150,000 s, 200,000 s, and 280,000 s (n = 3). Under gravity, molecules accumulate downward with molecular density decreasing exponentially with height. Since methane molecules exhibit higher density than hydrogen molecules, this results in a non-uniform spatial distribution characterized by a sparse upper and dense lower profile. At t = 280,000 s (3.24 days), hydrogen concentration at the top reaches 10.8%, representing a 0.8% increase from the initial state.
As shown in Figure 17, the hydrogen accumulation rate progressively decreases with standing time. After 280,000 s (3.24 days), the maximum hydrogen concentration at monitoring points reaches 10.65% (Figure 18).
When operational pressure and initial hydrogen concentration remain constant, the maximum hydrogen concentration at the top of the undulating pipeline exhibits a decreasing trend with an increasing number of spans after 3 days of standing (Table 10).
The gas mixture within the undulating pipeline does not reach a steady state even after 72 h of standing. Compared to horizontal pipelines, the stratification process requires significantly longer time. After 72 h, the maximum hydrogen concentration accumulated at the pipeline top increases with larger elevation differences but decreases with greater top span lengths or higher numbers of spans. However, the maximum hydrogen concentration remains close to the initial concentration, indicating insignificant stratification.

6. Conclusions

Computational fluid dynamics (CFD) models were developed to systematically analyze the flow dynamics and spatial distribution patterns of methane–hydrogen mixtures in both horizontal and undulating pipeline configurations. The simulations elucidate the coupled effects of hydrogen blending ratio and operational pressure on mixture stratification in horizontal pipelines, while further investigating the multifactorial interplay of blending ratio, pressure gradients, and elevation differentials in undulating pipeline systems. Key findings are summarized as follows:
(1)
In horizontal hydrogen-blended natural gas pipelines, progressive pressurization (from 3 MPa to 8 MPa) induced a transient stabilization phase, beyond which the equilibrium hydrogen concentration exhibited negligible deviation from the initial blending ratio.
(2)
In undulating pipelines, hydrogen concentration evolution at pipeline crests after a 72 h stagnation period exhibited the following patterns:
(i)
The data indicate that when operational conditions and pipeline geometry remain unchanged, increasing the elevation difference leads to higher maximum hydrogen concentrations at the pipeline top after 3 days of standing—rising from 3.74% at 10 m to 9.98% at 50 m.
(ii)
Under constant operational pressure and initial hydrogen concentration, the maximum hydrogen concentration exhibited a decreasing trend with increased crest span length. However, the peak concentration remained proximate to the initial value, indicating homogeneous methane–hydrogen mixing.
(iii)
Under fixed operational pressure and initial hydrogen concentration, the maximum hydrogen concentration remained proximal to the initial value with increasing span count, demonstrating enhanced homogeneity in methane–hydrogen mixtures.

Author Contributions

Conceptualization, T.L., J.X., H.Z. and Y.L.; Methodology, Y.L.; Validation, T.L., J.X. and H.Z.; Formal analysis, T.L.; Investigation, T.L., J.X. and H.Z.; Resources, J.C., K.L., Y.W. and Y.L.; Writing—original draft, T.L.; Writing—review & editing, Y.L.; Visualization, J.C., K.L. and Y.W.; Supervision, Y.L.; Project administration, Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key R&D Program (2022YFB4003402), and the Sichuan Science and Technology Program (No.2022NSFSC0028). Without their support, this work would not have been possible.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

Authors Tianlei Li, Ke Li and Yaxi Wang were employed by the company China Petroleum Engineering Construction Corporation. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Prajapati, A.; Sartape, R.; Rojas, T.; Dandu, N.K.; Dhakal, P.; Thorat, A.S.; Xie, J.; Bessa, I.; Galante, M.T.; Andrade, M.H. Migration-assisted, moisture gradient process for ultrafast, continuous CO2 capture from dilute sources at ambient conditions. Energy Environ. Sci. 2022, 15, 680–692. [Google Scholar] [CrossRef]
  2. Mazloomi, K.; Gomes, C. Hydrogen as an energy carrier: Prospects and challenges. Renew. Sustain. Energy Rev. 2012, 16, 3024–3033. [Google Scholar] [CrossRef]
  3. Singh, S.; Jain, S.; Venkateswaran, P.S.; Tiwari, A.K.; Nouni, M.R.; Pandey, J.K.; Goel, S. Hydrogen: A sustainable fuel for future of the transport sector. Renew. Sustain. Energy Rev. 2015, 51, 623–633. [Google Scholar] [CrossRef]
  4. Moradi, R.; Groth, K.M. Hydrogen storage and delivery: Review of the state of the art technologies and risk and reliability analysis. Int. J. Hydrogen Energy 2019, 44, 12254–12269. [Google Scholar] [CrossRef]
  5. Kojima, Y. Hydrogen storage materials for hydrogen and energy carriers. Int. J. Hydrogen Energy 2019, 44, 18179–18192. [Google Scholar] [CrossRef]
  6. Durbin, D.J.; Malardier-Jugroot, C. Review of hydrogen storage techniques for on board vehicle applications. Int. J. Hydrogen Energy 2013, 38, 14595–14617. [Google Scholar] [CrossRef]
  7. Ogden, J.; Jaffe, A.M.; Scheitrum, D.; McDonald, Z.; Miller, M. Natural gas as a bridge to hydrogen transportation fuel: Insights from the literature. Energy Policy 2018, 115, 317–329. [Google Scholar] [CrossRef]
  8. Dodds, P.E.; Staffell, I.; Hawkes, A.D.; Li, F.; Grunewald, P.; McDowall, W.; Ekins, P. Hydrogen and fuel cell technologies for heating: A review. Int. J. Hydrogen Energy 2015, 40, 2065–2083. [Google Scholar] [CrossRef]
  9. Chai, S.Q.; Zhang, G.J.; Li, G.Q.; Zhang, Y.F. Industrial hydrogen production technology and development status in China: A review. Clean Technol. Environ. Policy 2021, 23, 1931–1946. [Google Scholar] [CrossRef]
  10. Liu, W.; Wan, Y.M.; Xiong, Y.L.; Gao, P.B. Green hydrogen standard in China: Standard and evaluation of low-carbon hydrogen, clean hydrogen, and renewable hydrogen. Int. J. Hydrogen Energy 2022, 47, 24584–24591. [Google Scholar] [CrossRef]
  11. Ren, X.S.; Dong, L.C.; Xu, D.; Hu, B.S. Challenges towards hydrogen economy in China. Int. J. Hydrogen Energy 2020, 45, 34326–34345. [Google Scholar] [CrossRef]
  12. Meng, X.Y.; Gu, A.; Wu, X.G.; Zhou, L.L.; Zhou, J.; Liu, B.; Mao, Z.Q. Status quo of China hydrogen strategy in the field of transportation and international comparisons. Int. J. Hydrogen Energy 2021, 46, 28887–28899. [Google Scholar] [CrossRef]
  13. Li, Y.F.; Shi, X.P.; Phoumin, H. A strategic roadmap for large-scale green hydrogen demonstration and commercialisation in China: A review and survey analysis. Int. J. Hydrogen Energy 2022, 47, 24592–24609. [Google Scholar] [CrossRef]
  14. Deymi-Dashtebayaz, M.; Ebrahimi-Moghadam, A.; Pishbin, S.I.; Pourramezan, M. Investigating the effect of hydrogen injection on natural gas thermo-physical properties with various compositions. Energy 2019, 167, 235–245. [Google Scholar] [CrossRef]
  15. 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]
  16. Hairuddin, A.A.; Yusaf, T.; Wandel, A.P. A review of hydrogen and natural gas addition in diesel HCCI engines. Renew. Sustain. Energy Rev. 2014, 32, 739–761. [Google Scholar] [CrossRef]
  17. Park, C.; Kim, C.; Choi, Y. Power output characteristics of hydrogen-natural gas blend fuel engine at different compression ratios. Int. J. Hydrogen Energy 2012, 37, 8681–8687. [Google Scholar] [CrossRef]
  18. Sierens, R.; Rosseel, E. Variable composition hydrogen/natural gas mixtures for increased engine efficiency and decreased emissions. J. Eng. Gas Turbines Power 2000, 122, 135–140. [Google Scholar] [CrossRef]
  19. Dimopoulos, P.; Rechsteiner, C.; Soltic, P.; Laemmle, C.; Boulouchos, K. Increase of passenger car engine efficiency with low engine-out emissions using hydrogen–natural gas mixtures: A thermodynamic analysis. Int. J. Hydrogen Energy 2007, 32, 3073–3083. [Google Scholar] [CrossRef]
  20. Abohamzeh, E.; Salehi, F.; Sheikholeslami, M.; Abbassi, R.; Khan, F. Review of hydrogen safety during storage, transmission, and applications processes. J. Loss Prev. Process Ind. 2021, 72, 104569. [Google Scholar] [CrossRef]
  21. Yang, F.Y.; Wang, T.Z.; Deng, X.T.; Dang, J.; Huang, Z.Y.; Hu, S.; Li, Y.Y.; Ouyang, M.G. Review on hydrogen safety issues: Incident statistics, hydrogen diffusion, and detonation process. Int. J. Hydrogen Energy 2021, 46, 31467–31488. [Google Scholar] [CrossRef]
  22. Guo, L.J.; Su, J.Z.; Wang, Z.Q.; Shi, J.W.; Guan, X.J.; Cao, W.; Ou, Z.S. Hydrogen safety: An obstacle that must be overcome on the road towards future hydrogen economy. Int. J. Hydrogen Energy 2024, 51, 1055–1078. [Google Scholar] [CrossRef]
  23. Tong, L.L.; Peng, C.; Cao, X.W. One-dimensional model of hydrogen distribution in confined space based on Froude number. Ann. Nucl. Energy 2020, 138, 107210. [Google Scholar] [CrossRef]
  24. Peng, S.; He, Q.; Peng, D.; Ouyang, X.; Zhang, X.; Chai, C.; Zhang, L.; Sun, X.; Deng, H.; Hu, W.; et al. Equilibrium distribution and diffusion of mixed hydrogen-methane gas in gravity field. Fuel 2024, 358, 130193. [Google Scholar] [CrossRef]
  25. Pitts, W.M.; Yang, J.C.; Blais, M.; Joyce, A. Dispersion and burning behavior of hydrogen released in a full-scale residential garage in the presence and absence of conventional automobiles. Int. J. Hydrogen Energy 2012, 37, 17457–17469. [Google Scholar] [CrossRef]
  26. Khatiwada, D.; Vasudevan, R.A.; Santos, B.H. Decarbonization of natural gas systems in the EU–costs, barriers, and constraints of hydrogen production with a case study in Portugal. Renew. Sustain. Energy Rev. 2022, 168, 112775. [Google Scholar] [CrossRef]
  27. Gupta, R.; Guibentif, T.M.M.; Friedl, M.; Parra, D.; Patel, M.K. Macroeconomic analysis of a new green hydrogen industry using Input-Output analysis: The case of Switzerland. Energy Policy 2023, 183, 113768. [Google Scholar] [CrossRef]
  28. Strömer, S.; Beck, A.; Traninger, M.; Orsolits, D.; Reuter, S. Transitioning to a renewable hydrogen system: Optimal infrastructure for self-sufficient hydrogen supply in Austria by 2030. Smart Energy 2024, 15, 100151. [Google Scholar] [CrossRef]
  29. Simon, J.; Férriz, A.M.; Correas, L.C. HyUnder–hydrogen underground storage at large scale: Case study Spain. Energy Procedia 2015, 73, 136–144. [Google Scholar] [CrossRef]
  30. Borge-Diez, D.; Rosales-Asensio, E.; Icaza, D.; Açikkalp, E. The green hydrogen-water-food nexus: Analysis for Spain. Int. J. Hydrogen Energy 2024, 77, 1026–1042. [Google Scholar] [CrossRef]
  31. Andeobu, L.; Wibowo, S.; Grandhi, S. Renewable hydrogen for the energy transition in Australia—Current trends, challenges and future directions. Int. J. Hydrogen Energy 2024, 87, 1207–1223. [Google Scholar] [CrossRef]
  32. Sun, Y.H.; Ren, Y.N.; Cheng, Y.F. Dissociative adsorption of hydrogen and methane molecules at high-angle grain boundaries of pipeline steel studied by density functional theory modeling. Int. J. Hydrogen Energy 2022, 47, 41069–41086. [Google Scholar] [CrossRef]
  33. Azatyan, V.V.; Petukhov, V.A.; Prokopenko, V.M.; Timerbulatov, T.R. Possible Gravitational Stratification of Components in Mixtures of Reaction Gases. Russ. J. Phys. Chem. A 2019, 93, 986–987. [Google Scholar] [CrossRef]
  34. Badino, G. The Legend of Carbon Dioxide Heaviness. J. Cave Karst Stud. 2009, 71, 100–107. Available online: https://www.caves.org/wp-content/uploads/Publications/JCKS/v71/cave-71-01-100.pdf (accessed on 28 April 2025).
  35. Jayanthi, A.; Peddieson, J. Formulation of multiphase mixture models for fragmenting flows. Eur. J. Mech.-B/Fluids 2014, 46, 118–125. [Google Scholar] [CrossRef]
  36. Qi, Y.; Ishihara, T. Numerical study of turbulent flow fields around of a row of trees and an isolated building by using modified k-ε model and LES model. J. Wind. Eng. Ind. Aerodyn. 2018, 177, 293–305. [Google Scholar] [CrossRef]
  37. Lu, M.; Sun, X.H.; Li, Y.J.; Fan, Z.G. Turbulent numerical simulation method and its characteristic analysis. J. Hebei Univ. Archit. Technol. 2006, 2, 106–110. [Google Scholar]
  38. Launder, B.E.; Spalding, D.B. The numerical computation of turbulent flows. In Numerical Prediction of Flow, Heat Transfer, Turbulence and Combustion; Elsevier: Amsterdam, The Netherlands, 1983; pp. 96–116. [Google Scholar] [CrossRef]
  39. Ren, S.Y. Study on mixing and explosion propagation laws of natural gas in confined space. J. Saf. Sci. Technol. 2016, 12, 130–135. [Google Scholar]
Figure 1. Schematic of the simulated horizontal pipeline.
Figure 1. Schematic of the simulated horizontal pipeline.
Energies 18 03181 g001
Figure 2. Wavy Pipelines Model. L1 = bottom horizontal pipe length (m); L2 = top span length (m); H = elevation difference (m); D = pipe diameter (m).
Figure 2. Wavy Pipelines Model. L1 = bottom horizontal pipe length (m); L2 = top span length (m); H = elevation difference (m); D = pipe diameter (m).
Energies 18 03181 g002
Figure 3. Meshing of the horizontal pipeline.
Figure 3. Meshing of the horizontal pipeline.
Energies 18 03181 g003
Figure 4. Meshing of the undulating pipeline.
Figure 4. Meshing of the undulating pipeline.
Energies 18 03181 g004
Figure 5. After stabilization, comparison of experimental and simulation results for methane volume fractions at different positions inside the tank.
Figure 5. After stabilization, comparison of experimental and simulation results for methane volume fractions at different positions inside the tank.
Energies 18 03181 g005
Figure 6. Variation of hydrogen concentration along the vertical line in horizontal pipes with hydrogen concentrations of 3% (a), 5% (b), 10% (c), and 50% (d).
Figure 6. Variation of hydrogen concentration along the vertical line in horizontal pipes with hydrogen concentrations of 3% (a), 5% (b), 10% (c), and 50% (d).
Energies 18 03181 g006
Figure 7. Contour plot of hydrogen concentration variation over time at the z = 15 m cross-section in horizontal pipes with hydrogen concentrations of 3% (a), 5% (b), 10% (c), and 50% (d).
Figure 7. Contour plot of hydrogen concentration variation over time at the z = 15 m cross-section in horizontal pipes with hydrogen concentrations of 3% (a), 5% (b), 10% (c), and 50% (d).
Energies 18 03181 g007
Figure 8. Hydrogen concentration on the vertical cross-section of the undulating pipeline at different time points (elevation difference: 10 m).
Figure 8. Hydrogen concentration on the vertical cross-section of the undulating pipeline at different time points (elevation difference: 10 m).
Energies 18 03181 g008
Figure 9. Hydrogen concentration at cross-sections of the upper and lower horizontal segments in the undulating pipeline at different time points (elevation difference: 10 m).
Figure 9. Hydrogen concentration at cross-sections of the upper and lower horizontal segments in the undulating pipeline at different time points (elevation difference: 10 m).
Energies 18 03181 g009
Figure 10. Hydrogen concentration vs. time curves at six monitoring points (elevation difference: 10 m).
Figure 10. Hydrogen concentration vs. time curves at six monitoring points (elevation difference: 10 m).
Energies 18 03181 g010
Figure 11. Schematic of the undulating pipeline with 50 m top span length. “A-A” – “D-D” are hydrogen concentration monitoring surfaces, and “a” – “d” are hydrogen con-centration detection points.
Figure 11. Schematic of the undulating pipeline with 50 m top span length. “A-A” – “D-D” are hydrogen concentration monitoring surfaces, and “a” – “d” are hydrogen con-centration detection points.
Energies 18 03181 g011
Figure 12. Hydrogen concentration on the vertical cross-section at different time points (top span length: 50 m).
Figure 12. Hydrogen concentration on the vertical cross-section at different time points (top span length: 50 m).
Energies 18 03181 g012
Figure 13. Hydrogen concentration vs. time curves at four monitoring points (top span length: 50 m).
Figure 13. Hydrogen concentration vs. time curves at four monitoring points (top span length: 50 m).
Energies 18 03181 g013
Figure 14. Hydrogen concentration at cross-sections of the upper and lower horizontal segments at different time points (top span length: 50 m).
Figure 14. Hydrogen concentration at cross-sections of the upper and lower horizontal segments at different time points (top span length: 50 m).
Energies 18 03181 g014
Figure 15. Schematic of the undulating pipeline (n = 3). “A-A” – “G-G” are hydrogen concentration monitoring surfaces, and “a” – “g” are hydrogen concentration detection points.
Figure 15. Schematic of the undulating pipeline (n = 3). “A-A” – “G-G” are hydrogen concentration monitoring surfaces, and “a” – “g” are hydrogen concentration detection points.
Energies 18 03181 g015
Figure 16. Hydrogen concentration on the vertical cross-section at different time points (n = 3).
Figure 16. Hydrogen concentration on the vertical cross-section at different time points (n = 3).
Energies 18 03181 g016
Figure 17. Hydrogen concentration vs. time curves at monitoring points (n = 3).
Figure 17. Hydrogen concentration vs. time curves at monitoring points (n = 3).
Energies 18 03181 g017
Figure 18. Hydrogen concentration at cross-sections of the upper and lower horizontal segments at different time points. (n = 3).
Figure 18. Hydrogen concentration at cross-sections of the upper and lower horizontal segments at different time points. (n = 3).
Energies 18 03181 g018
Table 1. Hydrogen Pipelines in Different Regions of the World.
Table 1. Hydrogen Pipelines in Different Regions of the World.
RegionsPipeline Length
kmMiles
U.S.26081621
Europe1598993
Rest of World337209
World total45422823
Table 2. Hydrogen Pipelines of Different Global Companies.
Table 2. Hydrogen Pipelines of Different Global Companies.
CompaniesPipeline Length
kmMiles
Air Liquide19361203
Air Products1140708
Linde244152
Praxair739459
Others483300
World total45422823
Table 3. The concentration variation in the hydrogen and methane mixture at different heights.
Table 3. The concentration variation in the hydrogen and methane mixture at different heights.
Height (m)H2/CH4H2 PercentageCH4 Percentage
00.250020.00%80.00%
300.250420.03%79.97%
600.250820.05%79.95%
900.251320.08%79.92%
1200.251720.11%79.89%
10000.264220.90%79.10%
Table 4. Simulation parameters.
Table 4. Simulation parameters.
Pipeline TypeInternal Pressure (MPa)φH2,I (%)H (m)L2n
Horizontal pipeline33, 5, 10, 50---
53, 5, 10, 50---
83, 5, 10, 50---
Undulating pipeline8310, 30, 50101
8101053, 5
8101010, 50, 1001
Table 5. Grid independence verification.
Table 5. Grid independence verification.
Pipeline TypeMesh CountP1 Stabilized Hydrogen Concentration (%)Variation Rate (%)
Horizontal pipeline75,4623.56-
254,5663.703.93
446,2133.720.54
Undulating pipeline101,2453.63-
405,6123.743.03
620,1453.750.27
Table 6. Static stratification results for the horizontal pipe at 3 MPa internal pressure.
Table 6. Static stratification results for the horizontal pipe at 3 MPa internal pressure.
Internal Pressure (MPa)φH2, I (%)φH2, max (%)Stabilization Time (s)
333.3310,000
55.4310,000
1010.7310,000
5050.9010,000
Table 7. Static stratification results in horizontal pipelines at 5 MPa and 8 MPa internal pressures.
Table 7. Static stratification results in horizontal pipelines at 5 MPa and 8 MPa internal pressures.
Internal Pressure (MPa)φH2,I (%)φH2,max (%)Stabilization Time (s)
533.5010,000
55.7010,000
1011.0710,000
5051.1810,000
833.7010,000
55.7310,000
1011.5410,000
5051.610,000
Table 8. Static stratification simulation results for undulating pipelines with different elevation differences.
Table 8. Static stratification simulation results for undulating pipelines with different elevation differences.
Internal Pressure (MPa)L2 (m)nH (m)φH2,I (%)φH2,max
8 1011033.74%
306.63%
509.98%
Table 9. Comparison of Simulation Results for Settling Stratification in Wavy pipelines with Different L2.
Table 9. Comparison of Simulation Results for Settling Stratification in Wavy pipelines with Different L2.
Internal Pressure (MPa)L2 (m)nH (m)φH2,I (%)φH2,max
8 MPa10 m110 m1010.65%
50 m10.45%
100 m10.38%
Table 10. Comparison of Settling Stratification Simulation Results for Different Numbers of Spans.
Table 10. Comparison of Settling Stratification Simulation Results for Different Numbers of Spans.
Internal Pressure (MPa)L2 (m)nH (m)φH2,IφH2,max
8 MPa5 m310 m10%10.8%
510.78%
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

Li, T.; Xiao, J.; Zhang, H.; Cheng, J.; Li, K.; Wang, Y.; Lin, Y. Numerical Simulation Study of Gas Stratification in Hydrogen-Enriched Natural Gas Pipelines. Energies 2025, 18, 3181. https://doi.org/10.3390/en18123181

AMA Style

Li T, Xiao J, Zhang H, Cheng J, Li K, Wang Y, Lin Y. Numerical Simulation Study of Gas Stratification in Hydrogen-Enriched Natural Gas Pipelines. Energies. 2025; 18(12):3181. https://doi.org/10.3390/en18123181

Chicago/Turabian Style

Li, Tianlei, Jie Xiao, Honglin Zhang, Jinliang Cheng, Ke Li, Yaxi Wang, and Yuanhua Lin. 2025. "Numerical Simulation Study of Gas Stratification in Hydrogen-Enriched Natural Gas Pipelines" Energies 18, no. 12: 3181. https://doi.org/10.3390/en18123181

APA Style

Li, T., Xiao, J., Zhang, H., Cheng, J., Li, K., Wang, Y., & Lin, Y. (2025). Numerical Simulation Study of Gas Stratification in Hydrogen-Enriched Natural Gas Pipelines. Energies, 18(12), 3181. https://doi.org/10.3390/en18123181

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