1. Introduction
As an ideal energy source for the 21st century, hydrogen energy is abundant in resources and emits no carbon during its utilization. Its adoption holds profound significance for China in achieving carbon peak and carbon neutrality, addressing global climate change and building a community with a shared future for mankind [
1]. In recent years, to promote the transformation of China’s energy structure, reduce dependence on fossil fuels, and establish an environmentally friendly society, China has vigorously developed hydrogen energy. Many regions have incorporated the development of hydrogen energy into their 14th Five-Year Plans. Currently, the global demand for hydrogen is approximately 70 million tons per year (70 × 10
6 t/a), with a total market value of about 125 billion USD (125 × 10
9 USD).
The entire industry chain of hydrogen energy comprises three key stages: hydrogen production, hydrogen storage and transportation, and hydrogen utilization. Among these, large-scale, long-distance, and low-cost hydrogen storage and transportation represent the bottleneck restricting the development of hydrogen energy. To address this, some scholars have proposed utilizing existing natural gas pipelines for hydrogen transportation by blending hydrogen with natural gas. This approach avoids the challenges associated with newly constructed pure hydrogen pipelines, such as high initial investment costs, elevated operational expenses, and insufficient supporting infrastructure. Additionally, hydrogen can partially replace natural gas consumption, thereby reducing carbon emissions. Therefore, blending hydrogen into natural gas pipelines remains the optimal solution for hydrogen storage and transportation. However, the flow dynamics of hydrogen–natural gas mixtures in pipelines are not yet fully understood, making it imperative to conduct simulation research on the process of mixing hydrogen into natural gas pipelines.
Currently, natural gas–hydrogen blending pipeline projects remain limited globally but are undergoing a phase of rapid development. Selected operational and demonstration projects are summarized in
Table 1. As indicated, publicly disclosed projects are sparse, with critical operational parameters (e.g., pressure, blending ratios) often lacking transparency. Existing projects universally adopt hydrogen blending ratios below 25%, which aligns with the upper threshold (25%) set in this study. Nevertheless, these pioneering efforts underscore the growing momentum and promising potential of hydrogen–natural gas co-transportation technologies in decarbonizing energy systems.
In terms of physicochemical property variations in hydrogen-blended natural gas systems, Li et al. [
5] analyzed the impact of hydrogen on natural gas properties using the Soave–Benedict–Webb–Rubin (SBWR) equation of state (EOS). The study demonstrated that the isenthalpic curves rise with increasing hydrogen content, but the rate of ascent gradually diminishes. The Joule–Thomson coefficient was found to be more sensitive to hydrogen blending concentration at higher temperatures and lower pressures. Complementary research by Zhang Pu et al. [
6] demonstrated that at 40% HMR, key parameters including higher heating value (HHV), relative density, Wobbe index, viscosity, and volumetric energy density (at 10 MPa, 20 °C) experienced reductions of approximately 27%, 35%, 10%, 20%, and 40%, respectively. Concurrently, compressibility factor and sonic velocity showed increases of 18% and 34% under identical conditions. Notably, current natural gas specifications impose limitations on permissible HMR ranges, while existing analytical protocols and equipment hydrogen tolerance thresholds remain inadequate for precise characterization of hydrogen–natural gas mixtures. Fiebig et al. [
7] implemented the SmartSim pipeline network simulation tool to monitor gas quality parameters, particularly focusing on calorific value and compositional changes during hydrogen injection. In a separate investigation, Dell’Isola et al. [
8] computed thermodynamic properties and compressibility factors for 25% HMR mixtures, identifying significant alterations in relative density, heat capacity, and HHV, along with HMR-dependent increases in compressibility. Experimental studies by Kobayashi et al. [
9] utilized capillary viscometry quantified temperature/pressure-dependent viscosity characteristics of hydrogen–methane blends. Their measurements revealed non-linear viscosity reduction with increasing hydrogen content, with marked decreases becoming pronounced beyond 50% hydrogen concentration.
Regarding the safety of hydrogen-blended natural gas pipelines, Mahajan et al. [
10] demonstrated that hydrogen exhibits a higher propensity to penetrate pipeline materials or eventually leak due to its significantly smaller molecular size compared to methane. This characteristic results in hydrogen leakage rates that may be 1.3 to 2.8 times higher than those of methane. Furthermore, the combustion limits of methane and hydrogen differ considerably, with methane having a range of 4–15% [
11], while hydrogen exhibits a much broader range of 4–76% [
12]. Shirvill et al. [
13] reported that a hydrogen mole concentration below 25% does not generate significant explosion overpressure and thus does not substantially increase explosion risk. To assess the feasibility of hydrogen blending in natural gas pipelines, Cristello et al. [
14] concluded that a hydrogen mole concentration of less than 20% does not pose a safety threat to natural gas pipeline integrity. However, as the hydrogen mole concentration increases, the maximum allowable working pressure of the pipeline may exceed its original design limits. Hafsi et al. [
15] found that when the hydrogen mass ratio exceeded 2/3, the transient value of the maximum circumferential stress exceeded the allowable stress of pipeline steel X52. To prevent damage to the pipeline, they recommended limiting the hydrogen mass ratio to 30%. Subani et al. [
16] proposed a technology based on the transient pressure wave of hydrogen–natural gas mixture, which was used to detect and locate leaks in rigid pipelines. Their experimental results demonstrated a significant dependence of detection performance on the hydrogen mass ratio (HMR) in the gas mixture. Specifically, under sudden valve closure conditions, the study revealed that increasing HMR led to corresponding enhancements in three key parameters: transient pressure, wave celerity, and mass flux. This phenomenon substantiates the theoretical correlation between hydrogen concentration and the enhanced combustion characteristics of the gas mixture.
Regarding the flow behavior in pipelines after hydrogen blending, Liu Cuiwei et al. [
17] employed computational fluid dynamics (CFD) methods to demonstrate that at low flow rates, hydrogen tends to migrate toward the upper section of the pipeline, leading to the stratification of natural gas and hydrogen in hydrogen-blended natural gas pipelines. This stratification phenomenon is particularly pronounced under low-temperature and high-pressure conditions. Consequently, transporting hydrogen-blended natural gas at low pressure and high velocity is considered safer. Zhu Hongjun et al. [
18] investigated the static stratification phenomenon in hydrogen-blended natural gas pipelines with undulating terrain during pipeline shutdown. They found that greater terrain relief, longer pipeline lengths, and larger pipeline volumes result in higher hydrogen volume fractions required to achieve stable stratification in the top horizontal section of the pipeline, thereby prolonging the time required to reach stable stratification. Yan Shuangjie et al. [
19] proposed that the non-uniform mixing of natural gas and hydrogen is influenced by the composition of natural gas. Specifically, the higher concentrations of heavier hydrocarbons in natural gas, the longer the distance required for uniform mixing of natural gas and hydrogen. Zhang Heng et al. [
20] established a mathematical model for transporting hydrogen-blended natural gas and found that the frictional resistance loss in the pipeline after mixing hydrogen is reduced, and the volumetric flow rate of gas mixture is increased. Since the calorific value of hydrogen is only 1/3 of that of natural gas, the calorific value of hydrogen-doped natural gas is lower than that of pure natural gas. The increase in volumetric flow rate is insufficient to compensate for the reduction in calorific value, resulting in a decrease in the energy flow rate of the pipeline and a reduction in gas transmission efficiency. Yan Shuangjie et al. [
21] also constructed a three-dimensional model of a hydrogen- blended natural gas pipeline to study the effects of the hydrogen injection inlet structure and turbulence on the uniformity of natural gas and hydrogen mixing. The results showed that increasing the number of hydrogen injection inlets significantly reduces the distance required for uniform mixing of natural gas and hydrogen.
To predict the uniformity of hydrogen-blended natural gas, Suyue et al. [
22] established a deep neural network (DNN) model, utilizing the coefficient of variation (COV) of hydrogen concentration to characterize the uniformity of hydrogen mixing. The average error of COV predicted by this model was only 4.53%, and the computational efficiency was also two orders of magnitude faster than that simulated by CFD. Wang Shuai et al. [
23] established a T-junction hydrogen-blended natural gas pipeline. Their simulations using Simdroid and FLUENT revealed that increasing the hydrogen mixing ratio from 10% to 20% led to a significant rise in hydrogen concentration in the upper section of the pipeline, accompanied by more pronounced hydrogen stratification, with a width equivalent to one-third of the pipe diameter.
Yang Donghai et al. [
24] conducted a numerical investigation on five types of static mixers. The study revealed that the gas mixing uniformity increases with both velocity and hydrogen blending ratio. Concurrently, the intensification of turbulence induced by increased flow velocity enhances mass transfer and consequently improves mixing uniformity. Cadorin et al. [
25] employed Ansys CFX for finite element analysis of high-pressure gas flow in pipelines, discovering that hydrogen–methane mixtures (e.g., 90% methane and 10% hydrogen) under high Reynolds number conditions consume nearly twice as much energy as natural gas during transportation. This finding underscores the increased energy requirements for transporting hydrogen-blended natural gas. Tan et al. [
26] further demonstrated that the extent of energy transportation cost depends on the volume fraction of hydrogen and flow conditions. Their analysis revealed that the cost of transporting pure hydrogen is minimized when the inlet flow rate approximates that of pure methane transportation, while the cost peaks when hydrogen is transported at the same mass flow rate as methane.
Bainier et al. [
27] demonstrated that hydrogen injection into natural gas pipeline networks significantly reduces energy transmission efficiency. Their comparative analysis revealed that, under identical pressure ratios, the energy delivery decreased by 4%, 14%, and 15–20% for hydrogen mixing ratios of 10%, 40%, and 100%, respectively. Concurrently, the compressor energy input increased by 7%, 30%, and 210% for these respective ratios. This phenomenon is attributed to the combined effects of increased hydrogen content and high pressure drop, necessitating additional compressors to maintain energy supply within the gas network. In a related study, Cavana et al. [
28] developed a gas network model that demonstrated that hydrogen injection into natural gas distribution networks could effectively eliminate gas quality variations in residential lines without compromising existing transmission infrastructure. Quintino et al. [
29] further suggested that existing natural gas infrastructure could accommodate hydrogen with a volume fraction of 20% through minimal technical modifications. Gondal et al. [
30] investigated the hydrogen tolerance of various pipeline network components, revealing significant variations across equipment types. Their findings indicated that compressors, as critical components of natural gas transmission systems, are limited to a maximum hydrogen mixing ratio of 10%. Conversely, distribution networks and gas storage equipment can accommodate up to 50% hydrogen mixing, while end-user equipment typically accepts ratios between 20% and 50%.
Tabkhi et al. [
31] conducted a comprehensive assessment of hydrogen mixing impacts by setting the transmission power at 65% of the pipeline’s maximum capacity. Their analysis established a maximum permissible hydrogen mass ratio of 6.6%. Furthermore, they observed that maintaining constant energy flow in pipelines leads to increased energy consumption during transportation with higher hydrogen mixing ratios, necessitating additional compressor stations [
25,
27]. Wu et al. [
32] numerically investigated the mixing characteristics of hydrogen injected into pipelines at multiple angles. The study found that increasing the hydrogen blending ratio enhances the kinetic energy of hydrogen, thereby reducing its penetration process into methane and avoiding local enrichment. Despite these advancements, the transient flow characteristics of T-junction hydrogen-blended natural gas pipelines remain insufficiently explored, highlighting a critical research gap in this field.
In summary, the development of hydrogen-blended natural gas pipelines faces three primary challenges: the absence of standardized technical protocols, elevated safety risks, and prohibitive capital costs, all of which significantly impede the advancement of hydrogen transportation infrastructure. These challenges fundamentally stem from an incomplete understanding of methane–hydrogen mixing dynamics within pipeline systems. Addressing the core issue—designing safe and cost-effective hydrogen–natural gas pipelines while ensuring homogeneous mixing—requires comprehensive elucidation of the interplay between hydrogen blending ratios (HBRs) and turbulence intensity on mixture uniformity.
Heterogeneous hydrogen–methane distribution over extended distances or durations poses dual risks: (1) localized hydrogen partial pressure surges may accelerate material degradation, compromising pipeline integrity; (2) fluctuations in calorific value and flowrate of blended gas could undermine metering accuracy and optimal hydrogen injection positioning. To establish robust technical standards for hydrogen–natural gas systems, systematic investigations into these factors are imperative.
In this study, we employ numerical simulations to investigate the transient flow characteristics of hydrogen-blended natural gas in a T-junction configuration. Our methodology rigorously examines the temporal evolution of (i) molar fraction distributions, (ii) mixing homogeneity, and (iii) pressure drop dynamics. Furthermore, we systematically elucidate the effects of hydrogen injection rates and methane flow rates on these critical parameters. The quantitative analysis of biphasic mixing behavior presented in this study provides valuable insights and practical references for the engineering implementation of hydrogen-blended natural gas systems.
2. Materials and Methods
2.1. Fundamental Assumptions
The numerical simulation is performed using ANSYS Fluent 2023R2, a commercial CFD software package. To simplify the model and enhance computational efficiency while maintaining result accuracy, the following assumptions were made for the hydrogen mixing process in the T-junction hydrogen-blended natural gas pipeline investigated in this study:
- (1)
Natural gas is approximated as pure methane (CH4).
To validate this hypothesis, numerical simulations were performed using the natural gas composition reported by Uilhoorn [
33] (molar fractions: CH
4 = 98.3455%, C
2H
6 = 0.6104%, C
3H
8 = 0.1572%, n-C
4H
10 = 0.0299%, i-C
4H
10 = 0.0253%, n-C
5H
12= 0.0055%, i-C
5H
12= 0.0040%, N
2 = 0.0303%, CO
2 = 0.7918%) under hydrogen-blended pipeline conditions. The hydrogen molar concentration contours at the
Z = 0 cross-section were systematically compared with those obtained for pure methane gas under identical hydrogen blending scenarios, as shown in
Figure 1. The results demonstrate minimal discrepancies between the two gas compositions, indicating negligible sensitivity of hydrogen dispersion patterns to the inclusion of minor hydrocarbon impurities under the studied conditions. But for natural gas with very high heavy hydrocarbon content, the mixing effect of hydrogen and natural gas will slightly weaken [
21]. To rationalize the numerical modeling framework, this study exclusively considers natural gas with trace heavy hydrocarbon content. Consequently, the simplification of natural gas composition to pure methane represents a justified assumption, as the negligible concentration of heavy hydrocarbons (C
2+components) ensures minimal impact on the mixing phenomena under the investigated operational conditions.
- (2)
Both methane and hydrogen are treated as incompressible gases, maintaining constant density throughout the flow process.
The compressibility of a gas is typically assessed by calculating its Mach number (
Ma), where flows with
Ma < 0.3 are considered incompressible due to negligible density variations, while
Ma ≥ 0.3 necessitates compressibility considerations. Under the present study’s conditions—inlet temperature
T = 288 K, methane inlet velocity
v2 = 4 m/s, and specific heat ratio
γ ≈ 1.3 for the hydrogen–methane mixture—the Mach number is computed as follows:
From a Mach number perspective, gas compressibility effects can be considered negligible under the studied conditions (Ma < 0.3). However, this assumption necessitates further validation by analyzing the pressure gradient magnitude, as compressibility may still influence flow dynamics when local pressure variations exceed critical thresholds.
The compressibility criterion must also account for the relative pressure variation (ΔP/Pinlet). Compressibility-induced density changes become non-negligible when ΔP/Pinlet > 10%, even at low Mach numbers.
At elevated pipeline pressures, the following occurs: (1) Density-driven effects: Increased gas density amplifies density gradients, altering the balance between convective transport and molecular diffusion. Concurrently, higher Reynolds numbers (Re = ρvD/μ) enhance turbulent mixing through intensified eddy diffusion. (2) Diffusion suppression: According to the Chapman–Enskog theory, the binary diffusion coefficient decreases with pressure. (3) Shock-induced interfacial disruption: At high flow velocities (v→c, where c is the local speed of sound), steep pressure gradients may trigger shock waves, generating discontinuous density/temperature jumps that destabilize the hydrogen–methane interface and impede homogeneity.
To rigorously evaluate compressibility’s influence, comparative simulations were conducted with and without the density-based model under identical boundary conditions. Following methodologies from analogous studies on T-junction hydrogen–methane mixing (e.g., Khabbazi et al. [
34], Uilhoorn et al. [
33]), the density model employed the Soave–Redlich–Kwong (SRK) equation of state, while preserving all other geometric and flow parameters.
As shown in
Figure 2, the hydrogen molar fraction distributions exhibit negligible discrepancies between compressible and incompressible models at the investigated pressure range, with identical spatial patterns and quantitative values. To quantify this observation, the coefficient of variation (COV)—a uniformity metric detailed in
Section 2.10—was analyzed along the pipeline length (
Figure 3). A pink dashed box (denoted by the red arrow) highlights the locally enlarged area to provide detailed structural visualization. The variations in the coefficient of variation (COV) along the pipeline length are highly consistent. This also indicates that, under the conditions of this study (temperature of 288 K and pipeline pressure of 3.5 MPa), considering compressibility has little influence on the hydrogen–methane mixing process. This finding is also consistent with the results of Gondal [
30].
2.2. Governing Equations
The gas flow dynamics in a T-junction hydrogen-blended natural gas pipeline are governed by three fundamental conservation equations: the continuity equation (mass conservation), momentum conservation equation, and energy conservation equation. To accurately characterize the component distribution during hydrogen mixing and ensure the closure of the governing equation system, the species transport equation and turbulence model are additionally incorporated.
The selection of this turbulence model is justified by the similarity between the hydrogen injection process in the T-junction natural gas hydrogen-mixing pipeline and a circular jet flow. Kong Mingmin et al. [
35] confirmed that the simulation results are consistent with those generated by the realizable k-epsilon turbulence model. Furthermore, a comparative analysis of pressure drop values under varying flow velocities was conducted using three turbulence models: the Large Eddy Simulation (LES) with the WALE subgrid-scale model, the Realizable k-ε model, and the k-ω SST model. The simulated results were validated against theoretical pressure drop predictions, as illustrated in
Figure 4 (The theoretical pressure drop depicted in Figure is discussed in detail in
Section 2.9, where the underlying principles are analytically derived). The data reveal that all three models exhibit close agreement in pressure drop calculations, with maximum deviations of 18.143% observed for the k-ω SST model. While the LES model demonstrated the smallest overall error margin, the Realizable k-ε model was ultimately selected for numerical model implementation to optimize the balance between computational accuracy and transient simulation time costs.
Continuity equation (mass conservation equation):
where
represents the density of fluid;
t denotes the flow time; and
signifies the velocity in the direction
i-direction.
Momentum conservation equation:
where
p signifies the pressure and
Fj represents the mass force per unit mass acting in the
j-direction.
Energy conservation equation:
where
k represents the heat transfer coefficient of gas;
cp denotes the specific heat capacity of the gas mixture;
T represents the temperature; and
ST indicates the volumetric heat source term within the fluid.
Component transport equation:
where
Yi denotes the mass fraction of component
i;
Ri represents the net production rate of the chemical reaction;
Si signifies the production rate arising from discrete phase contributions and user-defined source terms; and Ji indicates the diffusion flux of component
i.
where
Di represents the diffusion coefficient.
Realizable k-epsilon turbulence equation:
where
k represents turbulence kinetic energy;
ε denotes the turbulent kinetic energy dissipation rate;
ut signifies the eddy viscosity coefficient;
YM indicates the contribution of expansion in compressible turbulence to the total dissipation rate;
Gk represents the turbulent kinetic energy of the average velocity;
Gb denotes the turbulent kinetic energy generated by buoyancy; C
1ε, C
2ε, C
3ε,
σk signifies the constant, respectively, taken as 1.22, 1.90, 0.09, 1.0;
σε indicates the Prandtl number of the turbulent kinetic energy dissipation rate, taken as 1.2;
represents the velocity vector;
xi signifies the position vector; and
Sij represents the strain rate tensor.
2.3. Geometric Model
The geometric configuration of the T-junction hydrogen-blended natural gas pipeline system is established in this study, as illustrated in
Figure 5. The system comprises a methane transmission main pipeline (
D = 80 mm) and a perpendicular hydrogen injection branch positioned 800 mm (10
D) downstream from the main inlet. The injection branch diameter was set at d = 0.2
D (16 mm) based on typical industrial blending ratios. Downstream of the injection point, the pipeline extends 8000 mm (100
D) to ensure a fully developed flow profile, incorporating five monitoring cross-sections spaced equidistantly at 1600 mm (20
D) intervals for boundary layer analysis. Methane was supplied through the upstream section while hydrogen was injected vertically via the side branch, enabling mixture formation governed by the Reynolds number (Re =
ρVD/
μ). The physical properties of methane and hydrogen transported by the T-junction natural gas hydrogen-mixing pipeline are summarized in
Table 2.
2.4. Boundary Conditions
The computational domain employs four boundary types: (a) mass flow inlet boundaries for methane supply (main pipe) and hydrogen injection (branch pipe), (b) pressure outlet boundary with outflow condition, (c) no-slip walls under isothermal conditions (
Tw = 288.15 K), and (d) Species transport model defining methane mass fraction as 1.0 at main inlet and 0.0 at the hydrogen branch. The methane flow rate values were selected based on the engineering specification for the design of a hydrogen transmission pipeline [
36], which recommends a maximum allowable flow velocity of 20 m/s for pipeline systems. However, considering the relatively small diameter of the simulated pipeline in this study, we adopted a more conservative maximum methane flow velocity of 10 m/s to ensure operational safety and alignment with empirical guidelines. Operating parameter settings are shown in
Table 3.
2.5. Mesh Strategy
The computational meshing methodology critically impacts numerical accuracy and convergence behavior. This study employs a hybrid Poly-Hexcore meshing based on “Mosaic” technology, strategically combining hexahedral-dominated core regions with polyhedral transitional zones. Three key meshing principles were implemented: (1) The Mosaic-enabled Poly-Hexcore topology achieves seamless node matching between hexahedral and polyhedral domains (As shown in the enlarged view above the pipeline in
Figure 6), eliminating manual interface treatment and reducing total cell count by 38% compared to conventional hex-dominant meshing; (2) Local mesh refinement at the T-junction confluence between the main pipe and the branch pipe (minimum cell size = 0.02
D, growth rate = 1.2), as shown in the enlarged view below the pipeline in
Figure 6; (3) The boundary layer mesh near the wall is also processed accordingly, and the scalable wall function is matched. This methodology neglects explicit resolution of the viscous sublayer and buffer layer, instead determining the first grid layer height based on an appropriate y+ value to ensure computational efficiency while preserving accuracy. Prism layers with (y
+ < 15) satisfy the viscous sublayer requirements of the realizable
k-
ε turbulence model, while maintaining orthogonal quality greater than 0.98 for 95% of cells. A near-wall treatment approach was employed to model the boundary layer, utilizing wall functions to approximate near-wall flow behavior.
To validate the rationality of the selected y+ range, first-layer grid height, and wall-function choice, simulations were conducted across seven operational conditions with mainstream velocities ranging from 4.44 m/s to 11.76 m/s. The first boundary layer grid height was set to 0.0013224776 m, corresponding to y+ values spanning 10–25 in the simulated results. This wide y+ range aligns with the applicability criteria of scalable wall functions, which are designed to handle moderate-to-high y+ regimes (typically y+ > 11) without requiring viscous sublayer resolution. The consistency of predicted y+ values across varying flow velocities confirms the robustness of the selected grid strategy and wall-function implementation. The boundary layer details are shown in
Figure 7.
2.6. Grid Independence Verification
Computational accuracy exhibits positive correlation with mesh density, while incurring significant computational resource demands. To optimize this trade-off, grid independence analysis was conducted with three systematically refined meshes: (1) fine mesh: 888,781 cells (baseline); (2) medium mesh: 752,910 cells (25% coarsening); and (3) coarse mesh: 617,172 cells (44% coarsening).
The outlet section velocity of the three meshes were statistically analyzed. The results showed that the relative error between the simulated values of the outlet average velocity was from 0.83% to 1.89%, respectively, from fine to coarse, which met the accuracy requirements.
Axial velocity profiles at four transverse planes (Y = 43.75
D, 56.25
D, 68.75
D, 81.25
D) were comparatively analyzed (
Figure 8). The medium-fine mesh comparison revealed less than 1.5% maximum velocity discrepancy, confirming spatial discretization adequacy. Consequently, the medium mesh was adopted for subsequent simulations, achieving 18.4% computational efficiency gain over the fine mesh while maintaining a less than 2% solution error.
2.7. Temporal Discretization Criteria
For transient numerical simulations, the selection of an optimal time step is critical to ensure numerical stability while minimizing computational cost. This study employs the Courant–Friedrichs–Lewy (CFL) condition to determine the maximum allowable time step:
where courant take is 0 < courant < 1; Δ
x denotes the grid spacing; and
u represents the fluid flow velocity.
This stability criterion essentially constrains the number of grid cells that fluid can traverse within a single time step. An excessively large time step would lead to insufficient spatial resolution to resolve critical flow features, whereas an overly small time step would impose prohibitive computational costs. In accordance with the FLUENT Official User’s Guide, which recommends a CFL value of 1 for transient simulations, and consistent with prior studies on gas pipeline modeling where CFL values ≤ 1 are standard practice [
37,
38,
39], we carefully balanced numerical stability and computational efficiency in this work. To ensure robustness while maintaining reasonable simulation times, a conservative CFL value of 0.8 was adopted. To balance numerical accuracy and computational efficiency, the time step in this study was dynamically determined through rigorous implementation of the CFL condition with a safety factor of 0.8.
2.8. Numerical Methodology
The computational framework fundamentally involves solving governing equations through iterative numerical schemes. Modern solvers provide multiple pressure–velocity coupling algorithms, including SIMPLE, SIMPLEC, PISO, and coupled methods. Strategic selection of these algorithms enhances convergence rates, optimizes computational efficiency, and prevents resource overutilization. Comparative studies evaluating SIMPLE, SIMPLEC, and PISO algorithms for lid-driven cavity flows demonstrate that while PISO exhibits superior robustness, it incurs higher computational overhead. Conversely, the SIMPLE algorithm achieves competitive convergence characteristics when paired with appropriate under-relaxation factors. Informed by these findings, the SIMPLE algorithm was employed in this study. The specific discretization strategies parameters are as follows:
Pressure: Second-order central differencing;
Momentum: Second-order upwind scheme;
Turbulent kinetic energy: First-order upwind scheme;
Turbulent dissipation rate: First-order upwind scheme;
Species transport convection term: Second-order upwind scheme;
Temporal discretization: First-order implicit formulation.
2.9. Model Validation
To validate the reliability of the T-junction hydrogen-blended natural gas pipeline model, the simulated pressure drop was benchmarked against theoretical predictions derived from the empirical pressure-loss coefficient correlation for T-junctions proposed by A. Gardel [
40]. The pressure-loss coefficient (
K32) for the T-junction (illustrated in
Figure 9) is calculated using Gardel’s formula [
40]:
where
,
.
Figure 4 illustrates the comparison between simulated and theoretical pressure drop values. The results demonstrate strong agreement between simulation outputs and theoretical calculations, with a maximum deviation of less than 4.5%. This confirms the validity and reliability of the adopted T-junction hydrogen–methane pipeline numerical model.
2.10. Evaluation of Mixing Uniformity
Various methodologies are currently available for assessing mixing uniformity. In this study, the coefficient of variation (COV), a dimensionless quantitative measure, was employed to evaluate the mixing homogeneity between methane and hydrogen. The COV is mathematically defined as the ratio of the standard deviation to the mean value of the sample volume fraction. The mathematical expression [
41,
42,
43,
44,
45] is presented below:
where
ci represents the volume fraction of hydrogen gas at the sampling point;
denotes the average volume fraction of all samples; and
n represents the total number of sample points.
The coefficient of variation (COV) ranges from 0 to 1. A higher COV value indicates greater non-uniformity in gas mixing, while a lower COV value corresponds to enhanced mixing uniformity. In engineering applications, a mixture is typically considered homogeneous when the COV is less than or equal to 0.05. Therefore, this criterion has been adopted in the present study.
Since the properties of hydrogen and methane calculated in this study closely approximate those of ideal gases, the mole fraction of hydrogen can be effectively equated to its volume fraction. Consequently, the volume fraction was calculated based on the mole fraction of hydrogen [
46].
To comprehensively investigate the hydrogen distribution, nine monitoring curves were established along the pipeline, extending from the hydrogen mixing point to the pipeline outlet section. Each cross-section of the pipeline was equipped with nine monitoring points, as illustrated in
Figure 10, which provides a schematic representation of the monitoring point distribution. The precise coordinates of the monitoring curves are detailed in
Table 4.
Based on the COV (coefficient of variation) values, this study proposes three quantitative parameters for assessing mixing homogeneity, which are defined as follows:
(1) Uniform mixing length: The minimum axial distance required for the first pipeline cross-section to achieve 95% mixing uniformity (COV ≤ 0.05). This parameter is determined by calculating the COV values along the entire pipeline length and plotting them as an XY-plot (abscissa: axial position Y along the flow direction; ordinate: COV values). The corresponding axial position at the first instance where COV ≤ 0.05 represents the uniform mixing length.
(2) Initial uniform mixing time: The time duration required for the first pipeline cross-section to reach 95% mixing uniformity. This is obtained by computing the temporal evolution of COV values along the pipeline and generating an XY-plot (abscissa: flow time; ordinate: COV values). The initial occurrence of COV ≤ 0.05 corresponds to the initial uniform mixing time.
(3) Overall mixing time: The temporal difference between outlet and initial mixing times. The outlet uniform mixing time is identified by analyzing the COV values at the pipeline outlet cross-section and determining the first instance of COV ≤ 0.05 in the XY-plot (abscissa: flow time; ordinate: COV values). The overall mixing time is the difference between outlet uniform mixing time and initial uniform mixing time.
These parameters provide a systematic framework for characterizing mixing performance in pipeline systems through time-resolved and spatially resolved COV analyses.
4. Conclusions
This study employs numerical simulation methodologies to conduct transient state simulations of hydrogen-blended natural gas pipelines with T-junction configurations. The research systematically investigates the temporal evolution of hydrogen mole concentration, coefficient of variation (COV), and pressure drop within the pipeline system. Furthermore, it establishes the quantitative relationships between hydrogen mixing rates (HMRs) and methane flow rates, and their impacts on hydrogen concentration distribution, uniform mixing length, and uniform mixing time. The findings provide valuable scientific insights and practical references for the engineering design and operational optimization of hydrogen–natural gas blending systems in industrial applications. The principal conclusions of this research are as follows:
(1) The gas stratification phenomenon in the main pipeline can be categorized into gravitational stratification and radial stratification. Gravitational stratification is primarily driven by the density disparity between methane and hydrogen, where methane exhibits higher density compared to hydrogen. Radial stratification arises from two key factors: when the hydrogen jet-to-methane flow rate ratio exceeds critical thresholds, and the hydrogen jet penetration depth increases significantly. These conditions induce vortex dynamics in the injection zone, causing hydrogen migration from lateral regions to the upper section of the pipeline, thereby creating radial concentration gradients.
(2) The results demonstrate that increasing hydrogen injection volume from 10% to 25% significantly reduces the uniform mixing length from 100.000 D to 20.875 D. This parametric study reveals that the HMR serves as the predominant control parameter governing mixing efficiency. Concurrently, increasing methane flow rate from 4 m/s to 10 m/s reduces uniform mixing length from 90.875 D to 71.500 D. Compared with the amount of HMR, the influence of the flow velocity on the mixing distance is weak, but it is easy to operate.
(3) The temporal evolution of the pressure drop reveals three distinct trends: an initial rapid increase, followed by an immediate decrease, and concluding with a gradual reduction until reaching stability. Both HMR and methane flow rate variations influence the pipeline flow dynamics, with increased HMRs and methane flow rates leading to elevated flow rates and corresponding pressure drops. However, the HMR exhibits limited impact on the flow rate, whereas methane flow rate modifications significantly affect the flow rate. Consequently, the pressure drop demonstrates greater sensitivity to variations in methane flow rate compared to hydrogen injection volume.
(4) The coefficient of variation (COV) demonstrates distinct trends along the pipeline length and over time. Along the axial distance, the COV initially decreases rapidly, followed by a slight decrease or marginal increase, eventually reaching a steady decline. Cross-sectionally, the COV exhibits a rapid temporal decrease, subsequently displaying a slight increase before stabilizing at a constant value.
(5) The adjustment of HMR and methane flow rate significantly impacts the mixing dynamics in the pipeline system. Specifically, increasing the HMR primarily accelerates the mixing process in the initial section of the pipeline, while demonstrating minimal effect on the uniform mixing time at the outlet section. Conversely, increasing the methane flow rate shows limited influence on the mixing efficiency in the initial pipeline section, but effectively reduces the mixing time at the outlet section with enhanced stability and reduced fluctuation. A fundamental trade-off in engineering applications is established: adjusting HMR effectively reduces uniform mixing length but increases overall uniform time, while modifying methane flow rate shortens overall uniform mixing time at the expense of increased uniform mixing length.