Next Article in Journal
A Novel Analytical Model for Structural Analysis of Long-Span Hybrid Cable-Stayed Suspension Bridges
Previous Article in Journal
Dynamic Characteristics and Periodic Stability Analysis of Rotor System with Squeeze Film Damper Under Base Motions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Analysis of Temperature Distribution During Charging Process of Vertically Installed Hydrogen Tanks

AI & Mechanical System Center, Institute for Advanced Engineering, Yongin-si 17180, Republic of Korea
*
Author to whom correspondence should be addressed.
Appl. Sci. 2025, 15(3), 1193; https://doi.org/10.3390/app15031193
Submission received: 31 October 2024 / Revised: 8 January 2025 / Accepted: 17 January 2025 / Published: 24 January 2025

Abstract

:
Recent advancements in hydrogen tank technology have favored the use of composites over metals for weight reduction, although at the cost of a narrower operating temperature range. Therefore, this study aims to establish safety standards for charging tube skids—large tanks positioned vertically within hydrogen transport vehicles—by examining the variations in temperature distribution under different charging environments through numerical analysis. A two-dimensional axisymmetric model was developed based on ANSYS Fluent to incorporate the effect of buoyancy by applying gravitational conditions along the axial direction. The study analyzed the impact of various charging and environmental conditions (charging time (0~240 min), inlet temperature, ambient temperature, initial temperature (−20 °C~40 °C), and initial pressure (5~20 bar)) on temperature distribution, ultimately deriving a regression equation to predict the tank’s maximum temperature. The findings indicate that the initial temperature has the most significant correlation with the tank’s temperature, followed by charging time and ambient temperature. Inlet temperature and initial pressure demonstrated minimal influence within the study’s scope. The derivation of predictive formulas for the maximum temperatures of the tank’s three regions, all showing an R2 value of 0.99 or higher, highlights the practicality of these results. This study’s insights are expected to contribute to further research aimed at optimizing charging conditions for tube skids and other hydrogen tanks.

1. Introduction

The environmental issues caused by carbon emissions have led to a concerted effort to mitigate greenhouse gas (GHG) emissions resulting from energy consumption. In particular, renewable energy technologies such as solar and wind power have demonstrated economic benefits through ongoing research and development [1]. Additionally, significant research has been directed toward carbon capture, utilization, and storage technologies to prevent the emission of GHGs produced by fossil fuel consumption into the atmosphere [2]. However, renewable energy sources face challenges in providing a stable electricity supply due to significant load fluctuations [3]. Moreover, integrating carbon capture equipment into transportation modes, such as automobiles, is impractical due to constraints on space and weight. Consequently, hydrogen fuel has emerged as a promising alternative to overcome the limitations of these carbon reduction technologies. Hydrogen produced through the electrolysis of water offers a medium to store energy generated from renewable sources and has the potential to replace conventional fossil fuels in the transportation sector, with only water as a byproduct after chemical reactions.
Currently, the prevalent hydrogen fuel technology involves compressing hydrogen to high pressures, approximately 350 to 700 bars, to address the issue of hydrogen gas’s low density [4]. This results in high pressure changes in hydrogen tanks during charging, compared to other gas fuels such as compressed natural gas [5], and leads to rapid temperature increases within the tank due to the enthalpy rise from gas compression during the filling process. Recent advancements in high-pressure hydrogen tanks, for instance, type IV tanks, have incorporated composites with a limited operational temperature range for weight reduction [4]. Thus, ensuring the safe usage of hydrogen tanks necessitates the establishment of specific charging and discharging conditions by analyzing the temperature distribution within hydrogen tanks under various charging scenarios.
Most previous studies on predicting temperature variations in hydrogen tanks have focused on fuel tanks for vehicles. For example, Yang [6] analytically estimated the temperature change during the hydrogen refueling process in hydrogen fuel-cell vehicles using ideal gas equations and applied the reduced Helmholtz free energy to correct the inaccuracies of these equations at high pressures.
Numerous studies have employed numerical methods to examine the temperature dynamics in hydrogen tanks designated for vehicle use, with a particular emphasis on rapid filling processes [7,8,9,10,11,12,13,14,15,16,17,18,19,20]. These investigations have explored the impact of various factors—such as state of charge, ambient temperature, charging duration, and tank geometry—on the temperature distribution within the tank throughout the charging procedure. Given that rapid filling typically occurs within a short period of a few minutes, some studies have modeled and analyzed the tank’s complete three-dimensional (3D) structure [7,8,9,10,11]. However, due to the extensive computational resources and time required to analyze compressible fluids, other research has limited the analysis to either half of the 3D geometry [12] or a simplified two-dimensional (2D) axisymmetric representation of the tank [13,14]. Notably, Melideo et al. [15] conducted a comparative analysis using both a 2D axisymmetric model and a full 3D model of a hydrogen tank. Their findings indicated that the results from the 2D and 3D models closely aligned during the charging and discharging phases, wherein hydrogen is injected into or released from the tank at high velocities, weakening the effects of buoyancy.
This study focuses on a specific application of hydrogen tanks in “tube trailers”, which are specialized vehicles designed for the transportation of high-pressure hydrogen gas from production sites to refueling stations. These trailers house multiple tanks, known as “tube skids”, vertically installed and interconnected to store and transport significant quantities of hydrogen, as depicted in Figure 1. This configuration offers enhanced flexibility in vehicle integration compared to traditional horizontal tube skids, and it facilitates rapid hydrogen release upwards from the vehicle in case of an emergency.
Given the considerable duration required to fully charge a vertically structured tank, the numerical analysis had to accommodate a broader time range than that typically associated with rapid filling. Additionally, the process of hydrogen introduction at the tank’s bottom at comparatively slow velocities suggested a pronounced effect of buoyancy on the internal flow patterns within the tank. Accordingly, this study developed a numerical analysis model based on the commercial software ANSYS Fluent. The model is designed as a 2D axisymmetric framework to effectively simulate the entire charging process of a single tube skid by incorporating the influence of gravity and buoyancy effects.
Utilizing this numerical model, the study established criteria for charging tube skids, analyzing the impact of five charging and environmental conditions on the temperature distribution within the tube skids. The analysis yielded temperature data, from which the sensitivity of each variable to temperature increases in the tank was determined. Subsequently, temperature prediction formulas for different regions within the tank were derived based on the numerical analysis results.

2. Numerical Analysis Model

2.1. Geometry of Analysis Model

To efficiently simulate the charging process of a tube skid, we simplified the analysis model. Although a typical trailer houses around 40 hydrogen tanks, as depicted in Figure 1, analyzing each tank simultaneously would be impractical. Therefore, we focused our numerical analysis on a single tank, based on the assumption that the slow flow velocity of hydrogen—resulting from the tanks’ parallel configuration and their substantial combined storage capacity—ensures uniform charging across all tanks, rendering the analysis of a single tank representative of the entire system. Furthermore, we neglected the influence of the connecting conduit, considering it inconsequential to the overall charging dynamics.
The tank was modeled in 2D axisymmetric geometry for efficient analysis. Figure 2 illustrates the geometry and dimensions of an individual tank, where the inlet’s complex structure was simplified for numerical analysis. The materials constituting the inlet, shell, and liner were aluminum, carbon fiber-reinforced polymer (CFRP), and polyamide (PA), respectively. The tank measured approximately 2.6 m in length and 0.55 m in diameter, with a total volume of 400 L. With the cylinder positioned vertically within a tube trailer, hydrogen was introduced via an inlet at the bottom of the cylinder.
Given the tank’s vertical orientation, buoyancy aligned with the tank’s axis. Furthermore, the tank’s geometry and the primary forces acting on the hydrogen exhibited symmetry relative to the central axis, enabling a 2D axisymmetric analysis to accurately account for buoyancy effects. Figure 3 shows the numerical analysis domain and the generated mesh representing the central cross section of the 3D tank geometry for 2D axisymmetric analysis. Comprising approximately 40,000 quadrilateral cells—determined optimal through grid testing—the grid’s average quality was assessed at 0.71.

2.2. Governing Equations of Analysis Model

Analyzing the hydrogen charging process comprehensively required simulating both the turbulent flow within the tank, driven by the momentum and buoyancy of the incoming hydrogen, and the resultant temperature increase and heat transfer due to hydrogen compression. Accordingly, this study employed a set of governing equations, including a continuity equation for compressible flow simulation, a momentum equation incorporating gravity’s effect, and a shear stress transport (SST) turbulence model to simulate turbulent flow. The adoption of the turbulence model aligned with its prevalent use in numerous prior investigations into the charging dynamics of high-pressure hydrogen [8,11,14]. Additionally, an energy equation was applied to model heat transfer within the tank. The computational analysis utilized the coupled algorithm within a pressure-based solver framework. The governing equations used in this study are as follows.
Continuity equation:
ρ t + · ρ u = 0
Momentum equation:
ρ u t + · ρ u u = ρ g P + · μ u
Energy equation:
ρ c p T t + · ρ c p T u = · k T
Transport equation for k of the kω SST model:
ρ κ t + · ρ k u = · Γ k u + G κ Y k
Transport equation for ω of the kω SST model:
ρ ω t + · ρ ω u = · Γ ω u + G ω Y ω + D ω

2.3. Properties and Boundary Conditions

Given the study’s focus on a tube skid designed to withstand a maximum charging pressure of 450 bars, the simulation of the hydrogen charging process had to incorporate hydrogen properties at pressures up to 450 bars. While the ideal gas law is commonly employed for its simplicity in calculating gas properties at low pressures, it fails to account for molecular attractions [21], rendering it inadequate for high-pressure applications. Consequently, the van der Waals equation, which considers both the size of molecules and intermolecular forces as shown in Equation (6), was developed. Moreover, various cubic equations of state have been modified from the van der Waals equation to enhance accuracy. Among these, the Redlich–Kwong model [22] has been utilized in prior studies to simulate the hydrogen charging process [5,7,8,9,10].
P = R T V b a / T V ( V + b )
Park [23] evaluated the accuracy of three state equations—Redlich–Kwong, Soave–Redlich–Kwong, and Peng–Robinson—in predicting hydrogen’s properties, finding that the Peng–Robinson equation exhibited a relative error of up to 5% against the National Institute of Standards and Technology (NIST) database. Although the Redlich–Kwong and Soave–Redlich–Kwong equations demonstrated smaller errors than the Peng–Robinson model, they still showed a relative error of up to 2%. While such discrepancies might be negligible in specific contexts, this study opted to utilize NIST’s property data to enhance analysis accuracy by minimizing inherent errors in these equations. Therefore, the NIST real gas model [24], as provided by ANSYS Fluent v21.2, was employed to calculate hydrogen’s thermodynamic properties based on the NIST database. This model, previously applied in studies [13,14] simulating the hydrogen tank charging process, was chosen to represent the properties of high-pressure hydrogen accurately.
Regarding the materials of the tank shell, ANSYS Fluent v21.2’s default properties were used for aluminum, while property values for CFRP and PA were specified, as listed in Table 1.
To assess the impact of environmental and operational conditions on the temperature distribution during charging, initial and boundary conditions were established, as outlined in Table 2. A total of five variables were considered to analyze their effect on the tank’s temperature rise. Seasonal variations, represented by summer (40 °C) and winter (−20 °C) conditions, informed the set values for inlet, ambient, and initial temperatures. In the initialization step of each case, both the fluid and solid regions were initialized to the initial temperature as shown in Table 2, and the pressure in the fluid region was specified as the initial pressure. The pressure of the inlet was regulated to increase linearly from the initial pressure to the maximum capacity of the tank (450 bar) during the charging time, and the temperature of the inlet stream was specified as the inlet temperature. It was assumed that convective heat transfer by the outside air continuously occurred on the outer wall of the tank, the temperature of the outside air was applied as the ambient temperature, and the convective heat transfer coefficient was set as 6 W/(m2·K) assuming minimum air flow. Moreover, walls in contact with different materials share the mesh geometry and mediate heat transfer under coupled conditions.

3. Analysis of Temperature Distribution Inside Tube Skid According to Operating and Environmental Conditions

3.1. Verification of Analysis Model

The numerical model used in this study incorporates the properties of high-pressure hydrogen and various governing equations, utilizing a 2D geometric representation of the hydrogen tank. To verify the model’s reliability, we used a study by Zheng et al. [13], which similarly employed a 2D axisymmetric model alongside NIST property data. The tank geometry in Zheng et al.’s study was approximated using the digitizer function within Origin software. Given that Zheng et al.’s tank was oriented horizontally, our analysis model was adjusted by omitting the gravity component from the numerical model utilized in our research.
In Zheng et al.’s experiment, hydrogen pressure exhibited an approximately linear increase; accordingly, we set the inlet pressure’s rise from approximately 5 to 64 MPa over 180 s, based on the graph from the verification experiment presented in their paper. Zheng et al. did not specify values for the convective heat transfer coefficient of the tank’s outer surface; consequently, we assumed a convective heat transfer coefficient of 6 W/(m2 K). The temperature rise of the tank predicted by our model was then compared with the temperature increase derived from the verification experiment graph in Zheng et al.’s paper, with results illustrated in Figure 4. Although an error margin of up to 6.6 °C was observed in the tank’s temperature rise—attributable to the reliance on estimated values from the graph—the overall trends between the two datasets were consistent, with an error of less than 1 °C at both the initial and final points. Because our study’s focus was on the tank’s maximum temperature at the end of charging, the error at the final charging point was considered a more critical evaluative criterion than the peak error observed throughout the charging process.

3.2. Analysis of Effect of Tube Skid Charging Time

The maximum charging rate for a tube skid is primarily governed by the compressor’s capacity to supply high-pressure hydrogen, whereas the minimum charging speed can be adjusted based on environmental conditions through a regulator. Given that tube skids are configured by interconnecting multiple tanks in parallel, hydrogen can be charged faster than the designated rate when only a few tanks are being filled. Therefore, examining the temperature distribution within the tube skid relative to charging duration is crucial for defining operational standards. This study conducted numerical analyses across four scenarios, ranging from 1 to 4 h, with a benchmark charging period of 2 h.
Figure 5 shows the temperature distribution across the central section of the tube skid following charging. Notably, the upper region exhibited higher temperatures compared to the lower region. This temperature discrepancy arose as the cooler hydrogen, introduced at the tube skid’s base, warmed up during compression and ascended due to buoyancy. The incoming hydrogen from the inlet ascended a certain distance propelled by its momentum before descending again due to density variations, resulting in unstable flow patterns near the skid’s bottom and a pronounced temperature gradient in the section where mixing occurred. Moreover, Zheng et al. [13] indicated that a temperature gradient could also be present along the axis of a horizontal tank if the ratio between its axial length and diameter was substantial.
Accordingly, the position within the tube skid significantly influences the disparity between the average temperature (mass-weighted average) and the maximum temperature (vertex maximum) of the hydrogen. Specifically, shorter charging intervals increase the temperature difference between the upper and lower sections due to a rapid pressure rise and diminished heat transfer duration. The difference between the peak and average temperatures was 20.7 °C for case 1 (the shortest charging duration) and 8.3 °C for case 4 (the longest charging duration).
Figure 6 illustrates the peak temperature within each region of the tube skid as a function of charging time, including the hydrogen’s maximum temperature inside the tube skid, the solid region’s (tank shell) peak temperature, and the outer surface’s maximum temperature. Overall, a prolonged charging period corresponded with a narrowing gap between the average internal hydrogen temperature and the outer surface’s temperature, decreasing from 19.9 to 8.5 °C, attributable to an extended heat transfer duration. The extended period also allowed for more effective heat dissipation to the external air, establishing an inverse relationship between the tube skid’s maximum temperature and the charging duration. Table 3 presents numerical data detailing the maximum temperatures upon charging completion.

3.3. Analysis of Effect of Inlet Temperature

The inlet temperature of hydrogen introduced into the tube skid is primarily influenced by external environmental conditions. However, this temperature may increase as the hydrogen is compressed and fed into the tube skid. In scenarios involving rapid filling, precooling measures are often implemented prior to hydrogen injection to mitigate excessive temperature rise within the hydrogen tank [15]. Thus, an investigation was conducted to examine the effect of the inlet temperature on the temperature distribution within the tube skid.
Figure 7 illustrates the temperature distribution across the central section of the tube skid following the charging process according to the inlet temperature. Contrary to the observations in Figure 6, where temperature varied throughout the tank depending on the case, alterations in the inlet temperature predominantly affected the temperature distribution in the tank’s lower section, with no significant impact on the higher-temperature region at the top of the tank. This phenomenon is attributed to the increased influence of buoyancy, a consequence of the low flow velocity and diminished momentum of hydrogen over extended charging periods. As the inlet temperature decreases, the density difference between the incoming hydrogen and that within the tank’s upper portion increases, suggesting that changes in inlet temperature primarily influence the temperature difference between the tank’s upper and lower sections without markedly affecting the upper region’s temperature. The regional maximum temperature analysis, as depicted in Figure 8 and detailed in Table 4, further corroborated that the tank’s maximum temperature varied by only 0.5 °C in response to different inlet temperatures.
Figure 8 also reveals a distinctive pattern in case 6 during the initial phase of charging on the tank’s outer surface, diverging from other cases. This anomaly arose because the peak surface temperature initially appeared in the aluminum section of the inlet, which was in direct contact with hydrogen at 40 °C. However, as charging advanced, the location of the highest surface temperature shifted to the upper part of the tube skid, aligning with the trends observed in other cases.

3.4. Analysis of Effect of Ambient Temperature

The thermal conductivity of the tube skid was notably low, attributable to its composite-based construction. In cases other than rapid filling scenarios, the tube skid is exposed to air for extended durations during the charging process, leading to an increase in heat dissipation to the surrounding environment. To investigate the influence of ambient temperature, an analysis was conducted while holding the convective heat transfer coefficient constant at 6 W/(m2 K) and varying the ambient temperature. The results of this analysis are presented in Figure 9 and Figure 10.
Figure 9 illustrates the effect of ambient temperature on the overall temperature distribution within the tube skid, subsequently affecting the maximum temperature by region, as depicted in Figure 10. The data from Figure 10 reveal that ambient temperature exerted minimal impact at the onset of charging when the hydrogen temperature experienced a rapid increase. However, the discrepancy in maximum temperature across different scenarios progressively widened after approximately 10 min from the start of charging, attributable to differences in heat transfer to the external air. These findings highlight the importance of considering seasonal variations when establishing charging rates for tube skids. The internal hydrogen’s maximum temperature exhibited fluctuations of up to 25 °C depending on the ambient temperature, with comprehensive details provided in Table 5.

3.5. Analysis of Effect of Initial Temperature

The initial temperature of the tube skid acted as an environmental variable because both the tube skid and the internal hydrogen, after prolonged exposure to outdoor air, tended to reach thermal equilibrium with the ambient temperature. However, the tube skid temperature changed when hydrogen was either injected into or released from it. Consequently, operations involving frequent charging or discharging could lead to notable variations in the initial temperature at subsequent recharging instances. The impact of the initial temperature was examined, with results presented in Figure 11 and Figure 12 and Table 6.
Given an initial charging pressure of 15 bars, as specified in Table 2, the hydrogen mass constituted less than 5% of the total capacity. Owing to the substantial heat retention by the solid shell region, fluctuations in the tank’s initial temperature significantly altered the tube skid’s overall temperature distribution, as depicted in Figure 11. Notably, a comparison between Figure 10 and Figure 12 reveals that, compared to the changes in ambient temperature, variations in the initial temperature exerted a more pronounced effect on the maximum temperature. While the ambient temperature change resulted in a maximum hydrogen temperature change of approximately 25 °C, this variation increased to approximately 36 °C with the change in initial temperature.
In Figure 12, the discontinuity observed in the maximum hydrogen temperature for case 10 is attributed to the rapid temperature change when hydrogen at 20 °C was introduced into the region initially set at −20 °C. Subsequently, the maximum temperature remained stable for approximately 15 min due to the heat capacity before resuming its increase as the enthalpy across the entire hydrogen region sufficiently rose.

3.6. Analysis of Effect of Initial Pressure

The residual hydrogen quantity and pressure within a tube skid following unloading may differ based on when the release is halted, implying that the initial pressure might vary for each subsequent charging cycle. Alterations in the initial pressure can influence the enthalpy and temperature of the internal hydrogen at the end of charging, as they affect the compression ratio needed to reach the maximum pressure. Therefore, an analysis was conducted to determine the impact of initial pressure variations, with the assumption that the tank was nearly depleted to set a benchmark for maximum discharge. The investigation focused on observing any significant temperature disparities resulting from initial pressure adjustments within the 5 to 20 bars range.
According to Figure 13, which illustrates the central section’s temperature distribution upon charging completion, variations in the initial pressure from 5 to 20 bars did not significantly affect the final temperature distribution, with discrepancies under 4 °C. This phenomenon likely occurred because a lower initial pressure necessitated more compression as the tank filled, yet the relevance of compressing the initially present hydrogen diminished as its initial mass decreased. As shown in the case 13 graph in Figure 14a, the temperature initially surged when starting at a lower pressure, but then the rate of increase moderated, aligning with other cases as more hydrogen was introduced. Consequently, Figure 14 confirms a marginal uptrend in maximum temperature with decreasing initial pressure, although with relatively low sensitivity to the pressure changes. Detailed results are provided in Table 7.

4. Sensitivity Analysis for Factors and Temperature Relationship Derivation

Following the analysis of the five variables critical to the charging characteristics of vertically installed hydrogen tanks, a design of experiments approach was employed to derive equations predicting the maximum temperature within the internal hydrogen, shell, and outer surface of the tank. Initially, key factors influencing the tank’s maximum temperature were identified through sensitivity analysis using a factorial design. Subsequently, equations for predicting these maximum temperatures were formulated via response surface methodology.
Given the varied materials and permissible temperature ranges across different sections of the tube skid, the tank was segmented into hydrogen, shell, and outer surface regions for analysis. The validity of the charging variables for each region’s maximum temperature was assessed through factorial design. The charging time and temperature ranges were maintained at 60 to 240 min and −20 to 40 °C, respectively, consistent with the previous sections. The experimental combinations derived from factorial design are listed in Table 8 and Table 9, with analysis results presented in Table 10, Table 11 and Table 12. Adj SS represents the temperature variance analyzed statistically, while Adj MS is derived by dividing Adj SS by the degrees of freedom (DF). The F-value correlates with the ratio of each variable’s Adj MS to the error’s Adj MS, and the p-value inversely relates to the F-value. A p-value < 0.05 generally indicates a factor’s significance. Applying this criterion, charging time, ambient temperature, and initial temperature were identified as significant factors, whereas inlet temperature and initial pressure were considered nonsignificant, as evidenced in the tables. The analysis results are detailed in Figure 15.
Further, a response surface methodology analysis was conducted to establish a regression equation for predicting the tank’s maximum temperature using the three significant factors identified. A face-centered design was utilized, with data from an additional 15 cases analyzed, as shown in Table 13.
This design facilitated the derivation of the regression equation (Equation (7)) from the analysis results. In the regression equation, the validity of the terms can be identified by applying the reference value of the p-value (0.05) to each term, and the terms with no influence can be excluded from the equation. Here, Xi represents the design factors, Z denotes the result, and βi and βij are the coefficients for each term.
Z = β 0 + Σ i = 1 k β i X i + Σ i = 1 , j i k β i j X i X j
The response surface methodology’s analysis results for the tube skid are summarized in Table 14, Table 15 and Table 16, with Figure 16 illustrating the correlations between each variable and the tank’s maximum temperature. The charging time is represented by tchr, the ambient temperature by Tamb, and the initial temperature by Tini. By isolating only the impactful terms, the regression equation for the internal hydrogen’s maximum temperature is expressed as Equation (8), with a coefficient of determination (R2) of approximately 0.9977—surpassing the often-referenced benchmark of 0.8 [25]. This high R2 value suggests that the regression equation reliably predicted the maximum temperature, facilitating the establishment of charging standards for tube skids without necessitating further analysis.
Similarly, regression equations for the shell and outer surface temperatures were derived, resulting in Equations (9) and (10). The R2 values for these equations exceeded 0.99, confirming their accuracy in capturing the analytical trends.
T g a s = 96.69 0.3956 · t c h r + 0.073 · T a m b + 0.9605 · T i n i + 0.000722 · t c h r 2 + 0.002501 · t c h r · T a m b 0.002692 · t c h r · T i n i
T s h e l l = 90.63 0.3586 · t c h r + 0.0934 · T a m b + 0.9330 · T i n i + 0.000647 · t c h r 2 + 0.002413 · t c h r · T a m b 0.002578 · t c h r · T i n i
T s u r f a c e = 71.047 0.2381 · t c h r + 0.2023 · T a m b + 0.791 · T i n i + 0.000391 · t c h r 2   + 0.001269 · T i n i 2 + 0.00198 · t c h r · T a m b 0.002138 · t c h r · T i n i

5. Conclusions

To evaluate the effects of charging conditions on the temperature distribution within a 400L-class type IV hydrogen tank, referred to as tube skid, vertically positioned in a hydrogen transport vehicle, a 2D axisymmetric model incorporating the effect of buoyancy was developed. Comprehensive numerical analyses were performed, utilizing property data from NIST. The study examined the influence of five charging parameters (charging time, inlet temperature, ambient temperature, initial temperature, and initial pressure) on the tube skid’s maximum temperature, yielding the following insights:
  • The inlet temperature predominantly affected the temperature at the lower section of the tank due to buoyancy, as hydrogen was introduced from the bottom. This condition increased the temperature disparity between the tank’s upper and lower sections without significantly altering the maximum temperature at the top of the tank.
  • The initial pressure exhibited minimal impact on the maximum temperature, as the initial mass of hydrogen was relatively small within the examined pressure range of 5 to 20 bars.
  • Sensitivity analysis through factorial design further confirmed the exclusion of inlet temperature and initial pressure as significant factors. Consequently, three out of the five analyzed factors were identified as closely associated with the tank’s maximum temperature. The initial temperature exhibited the strongest correlation, followed by charging time and ambient temperature, highlighting the importance of these three parameters in the hydrogen charging process.
  • Utilizing the three identified critical factors, a regression equation for predicting the tank’s maximum temperature upon charging completion was formulated via response surface methodology. With the derived equation exhibiting an R2 value of 0.99 or higher, it facilitated highly accurate predictions of temperature across different regions of the tank using these parameters.
Thus, the findings from this study are poised to contribute to the development of charging standards for vertically installed tube skids and may inform subsequent research aimed at optimizing charging conditions.

Author Contributions

Conceptualization, Y.L., J.-H.N. and D.-J.H.; Methodology, Y.L. and J.-H.N.; Software, Y.L. and J.-H.N.; Data Analysis, Y.L. and J.-H.N.; Investigation, Y.L.; Writing—Original Draft Preparation, Y.L. and J.-H.N.; Writing—Review and Editing Y.L., J.-H.N. and D.-J.H.; Visualization, Y.L. and J.-H.N.; Supervision, J.-H.N. and D.-J.H.; Project Administration, J.-H.N.; Funding Acquisition, J.-H.N. and D.-J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Technology Innovation Program (20026413, Development of the BOP and system for PEM water electrolysis) funded By the Ministry of Trade Industry and Energy (MOTIE, Republic of Korea).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

Cpspecific heat, J kg−1 K−1
Dωcross-diffusion term, kg m−3 s−2
Gkproduction of turbulence kinetic energy, kg m−1 s−3
Gωgeneration of specific dissipation rate, kg m−3 s−2
Ppressure, Pa
Rgas constant, J K−1 mol−1
Ttemperature, K
Tiniinitial temperature, K
Tambambient temperature, K
Vmolar volume, m3 mol−1
X design factor for regression equation
Ykdissipation of turbulence kinetic energy, kg m−1 s−3
Yωdissipation of specific dissipation rate, kg m−3 s−2
Z result of regression equation
a   constant that corrects for attractive potential of molecules
b   constant that corrects for volume of molecules
g   gravity, m s−2
k turbulence kinetic energy, m2 s−2
k c thermal conductivity, W m−1 k−1
t time, s
tchrcharging time, s
u velocity, m s−1
β coefficient for each terms of regression equation
Γ k effective diffusivity for the turbulence kinetic energy, kg m−1 s−1
Γ ω effective diffusivity for the specific dissipation rate, kg m−1 s−1
γ heat capacity ratio
μ viscosity, kg m−1 s−1
ρ   density, kg m−3
ω specific dissipation rate, m2 s−3

Abbreviations

2 D two-dimensional
3 D three-dimensional
C F R P carbon fiber-reinforced polymer
G H G greenhouse gas
N I S T National Institute of Standards and Technology
P A Polyamide
S S T shear stress transport

References

  1. Sohag, K.; Chukavina, K.; Samargandi, N. Renewable energy and total factor productivity in OECD member countries. J. Clean. Prod. 2021, 296, 126499. [Google Scholar] [CrossRef]
  2. Yadav, S.; Mondal, S.S. A review on the progress and prospects of oxy-fuel carbon capture and sequestration (CCS) technology. Fuel 2022, 308, 122057. [Google Scholar] [CrossRef]
  3. Brauns, J.; Turek, T. Alkaline water electrolysis powered by renewable energy: A review. Processes 2020, 8, 248. [Google Scholar] [CrossRef]
  4. de Miguel, N.; Acosta, B.; Baraldi, D.; Melideo, R.; Cebolla, R.O.; Moretto, P. The role of initial tank temperature on refuelling of on-board hydrogen tanks. Int. J. Hydrogen Energy 2016, 41, 8606–8615. [Google Scholar] [CrossRef]
  5. Berghmans, J.; Vanierschot, M. Safety aspects of CNG cars. Procedia Eng. 2014, 84, 33–46. [Google Scholar] [CrossRef]
  6. Yang, J.C. A thermodynamic analysis of refueling of a hydrogen tank. Int. J. Hydrogen Energy 2009, 34, 6712–6721. [Google Scholar] [CrossRef]
  7. Kim, S.C.; Lee, S.H.; Yoon, K.B. Thermal characteristics during hydrogen fueling process of type IV cylinder. Int. J. Hydrogen Energy 2010, 35, 6830–6835. [Google Scholar] [CrossRef]
  8. Heitsch, M.; Baraldi, D.; Moretto, P. Numerical investigations on the fast filling of hydrogen tanks. Int. J. Hydrogen Energy 2011, 36, 2606–2612. [Google Scholar] [CrossRef]
  9. Galassi, M.C.; Baraldi, D.; Iborra, B.A.; Moretto, P. CFD analysis of fast filling scenarios for 70 MPa hydrogen type IV tanks. Int. J. Hydrogen Energy 2012, 37, 6886–6892. [Google Scholar] [CrossRef]
  10. Galassi, M.C.; Papanikolaou, E.; Heitsch, M.; Baraldi, D.; Iborra, B.A.; Moretto, P. Assessment of CFD models for hydrogen fast filling simulations. Int. J. Hydrogen Energy 2014, 39, 6252–6260. [Google Scholar] [CrossRef]
  11. Suryan, A.; Kim, H.D.; Setoguchi, T. Three dimensional numerical computations on the fast filling of a hydrogen tank under different conditions. Int. J. Hydrogen Energy 2012, 37, 7600–7611. [Google Scholar] [CrossRef]
  12. Melideo, D.; Baraldi, D.; Acosta-Iborra, B.; Cebolla, R.O.; Moretto, P. CFD simulations of filling and emptying of hydrogen tanks. Int. J. Hydrogen Energy 2017, 42, 7304–7313. [Google Scholar] [CrossRef]
  13. Zheng, J.; Guo, J.; Yang, J.; Zhao, Y.; Zhao, L.; Pan, X.; Ma, J.; Zhang, L. Experimental and numerical study on temperature rise within a 70 MPa type III cylinder during fast refueling. Int. J. Hydrogen Energy 2013, 38, 10956–10962. [Google Scholar] [CrossRef]
  14. Liszka, M.; Fridlyand, A.; Jayaraman, A.; Bonnema, M.; Sishtla, C. CFD modeling of the hydrogen fast filling process for type 3 cylinders and cylinders lined with phase change material. ASME Int. Mech. Eng. Congr. Expo. 2019, 59445, V007T08A061. [Google Scholar]
  15. Melideo, D.; Baraldi, D.; Galassi, M.C.; Cebolla, R.O.; Iborra, B.A.; Moretto, P. CFD model performance benchmark of fast filling simulations of hydrogen tanks with pre-cooling. Int. J. Hydrogen Energy 2014, 39, 4389–4395. [Google Scholar] [CrossRef]
  16. Liu, J.; Zheng, S.; Zhang, Z.; Zheng, J.; Zhao, Y. Numerical study on the fast filling of on-bus gaseous hydrogen storage cylinder. Int. J. Hydrogen Energy 2020, 45, 9241–9251. [Google Scholar] [CrossRef]
  17. Liu, Y.; Zheng, J.; Xu, P.; Zhao, Y.; Li, L.; Liu, P. Numerical simulation on fast filling of hydrogen for composite storage cylinders. In Proceedings of the ASME Pressure Vessels and Piping Division Conference, Chicago, IL, USA, 27–31 July 2008. [Google Scholar]
  18. Dmael, E.; Weber, M.; Renault, P. Experimental and numerical evaluation of transient temperature distribution inside a cylinder during fast filling for H2 applications. In Proceedings of the Eighteenth World Hydrogen Energy Conference, Essen, Germany, 16–20 May 2010. [Google Scholar]
  19. Takagi, Y.; Sugie, N.; Takeda, K.; Okano, Y.; Eguchi, T.; Hirota, K. Numerical investigation of the thermal behavior in a hydrogen tank during fast filling process. In Proceedings of the ASME/JSME 2011 8th Thermal Engineering Joint Conference, Honolulu, HI, USA, 13–17 March 2011. [Google Scholar]
  20. Galassi, M.C.; Acosta-Iborra, B.; Baraldi, D.; Bonato, C.; Harskamp, F.; Frischauf, N.; Moretto, P. Onboard compressed hydrogen storage: Fast filling experiments and simulations. Energy Procedia 2012, 29, 192–200. [Google Scholar] [CrossRef]
  21. Berberan-Santos, M.N.; Bodunov, E.N.; Pogliani, L. The van der Waals equation: Analytical and approximate solutions. J. Math. Chem. 2008, 43, 1437–1457. [Google Scholar] [CrossRef]
  22. Redlich, O.; Kwong, J.N. On the thermodynamics of solutions, V. An equation of state. Fugacities of gaseous solutions. Chem. Rev. 1949, 44, 233–244. [Google Scholar] [CrossRef]
  23. Park, B.H. Calculation and Comparison of Thermodynamic Properties of Hydrogen Using Equations of State for Compressed Hydrogen Storage. Trans. Korean Hydrog. New Energy Soc. 2020, 31, 184–193. [Google Scholar] [CrossRef]
  24. ANSYS Inc. ANSYS Fluent User’s Guide (Release 15.0); ANSYS Inc.: Canonsburg, PA, USA, 2013; pp. 487–492. [Google Scholar]
  25. Joglekar, A.M.; May, A.T. Product excellence through design of experiments. Cereal Foods World 1987, 32, 857. [Google Scholar]
Figure 1. Hydrogen transport system (tube trailer).
Figure 1. Hydrogen transport system (tube trailer).
Applsci 15 01193 g001
Figure 2. Geometry of the tube skid.
Figure 2. Geometry of the tube skid.
Applsci 15 01193 g002
Figure 3. Computational domain of the tube skid.
Figure 3. Computational domain of the tube skid.
Applsci 15 01193 g003
Figure 4. Validation of numerical model with literature data.
Figure 4. Validation of numerical model with literature data.
Applsci 15 01193 g004
Figure 5. Temperature distribution of central section at completion of charging by change in charging time.
Figure 5. Temperature distribution of central section at completion of charging by change in charging time.
Applsci 15 01193 g005
Figure 6. Effect of charging time on maximum temperature for each region of tube skid; (a) hydrogen, (b) tank shell, and (c) outer surface.
Figure 6. Effect of charging time on maximum temperature for each region of tube skid; (a) hydrogen, (b) tank shell, and (c) outer surface.
Applsci 15 01193 g006aApplsci 15 01193 g006b
Figure 7. Temperature distribution of central section at completion of charging by change in inlet temperature.
Figure 7. Temperature distribution of central section at completion of charging by change in inlet temperature.
Applsci 15 01193 g007
Figure 8. Effect of inlet temperature on maximum temperature for each region of tube skid; (a) hydrogen, (b) tank shell, and (c) outer surface.
Figure 8. Effect of inlet temperature on maximum temperature for each region of tube skid; (a) hydrogen, (b) tank shell, and (c) outer surface.
Applsci 15 01193 g008aApplsci 15 01193 g008b
Figure 9. Temperature distribution of central section at completion of charging by change in ambient temperature.
Figure 9. Temperature distribution of central section at completion of charging by change in ambient temperature.
Applsci 15 01193 g009
Figure 10. Effect of ambient temperature on maximum temperature for each region of tube skid; (a) hydrogen, (b) tank shell, and (c) outer surface.
Figure 10. Effect of ambient temperature on maximum temperature for each region of tube skid; (a) hydrogen, (b) tank shell, and (c) outer surface.
Applsci 15 01193 g010aApplsci 15 01193 g010b
Figure 11. Temperature distribution of central section at completion of charging by change in initial temperature.
Figure 11. Temperature distribution of central section at completion of charging by change in initial temperature.
Applsci 15 01193 g011
Figure 12. Effect of initial temperature on maximum temperature for each region of tube skid; (a) hydrogen, (b) tank shell, and (c) outer surface.
Figure 12. Effect of initial temperature on maximum temperature for each region of tube skid; (a) hydrogen, (b) tank shell, and (c) outer surface.
Applsci 15 01193 g012aApplsci 15 01193 g012b
Figure 13. Temperature distribution of central section at completion of charging by change in initial pressure.
Figure 13. Temperature distribution of central section at completion of charging by change in initial pressure.
Applsci 15 01193 g013
Figure 14. Effect of initial pressure on maximum temperature for each region of tube skid; (a) hydrogen, (b) tank shell, and (c) outer surface.
Figure 14. Effect of initial pressure on maximum temperature for each region of tube skid; (a) hydrogen, (b) tank shell, and (c) outer surface.
Applsci 15 01193 g014aApplsci 15 01193 g014b
Figure 15. Responsivities of performance factors using fractional factorial design.
Figure 15. Responsivities of performance factors using fractional factorial design.
Applsci 15 01193 g015aApplsci 15 01193 g015b
Figure 16. Main effects plot for maximum temperature of hydrogen tube skid.
Figure 16. Main effects plot for maximum temperature of hydrogen tube skid.
Applsci 15 01193 g016aApplsci 15 01193 g016b
Table 1. Properties of tube skid materials.
Table 1. Properties of tube skid materials.
PropertiesAl6061PACFRP
Density [kg/m3]271911001540
Specific Heat [J/(kg K)]89618721040
Thermal Conductivity [W/(m K)]1670.3341.2
Table 2. Numerical analysis list for operating conditions and environmental conditions.
Table 2. Numerical analysis list for operating conditions and environmental conditions.
CaseCharging TimeInlet Temperature (°C)Ambient Temperature (°C)Initial Temperature (°C)Initial Pressure (bar)
Ref2 h (120 min)20202015
11 h (60 min)20202015
23 h (180 min)20202015
34 h (240 min)20202015
42 h (120 min)−20202015
52 h (120 min)0202015
62 h (120 min)40202015
72 h (120 min)20−202015
82 h (120 min)2002015
92 h (120 min)20402015
102 h (120 min)2020−2015
112 h (120 min)2020015
122 h (120 min)20204015
132 h (120 min)2020205
142 h (120 min)20202010
152 h (120 min)20202020
Table 3. Maximum temperature for each region of tube skid by charging time.
Table 3. Maximum temperature for each region of tube skid by charging time.
CaseCharging TimeMaximum Temperature
(Hydrogen, °C)
Maximum Temperature
(Tank Shell, °C)
Maximum Temperature
(Outer Surface, °C)
11 h (3600 s)96.191.779.0
Ref2 h (7200 s)79.176.467.7
23 h (10,800 s)69.667.661.2
34 h (14,400 s)62.861.356.3
Table 4. Maximum temperature for each region of tube skid by inlet temperature.
Table 4. Maximum temperature for each region of tube skid by inlet temperature.
CaseInlet TemperatureMaximum Temperature
(Hydrogen, °C)
Maximum Temperature
(Tank Shell, °C)
Maximum Temperature
(Outer Surface, °C)
4−20 °C79.176.367.7
50 °C79.276.567.9
Ref20 °C79.176.467.7
640 °C79.676.868.1
Table 5. Maximum temperature for each region of tube skid by ambient temperature.
Table 5. Maximum temperature for each region of tube skid by ambient temperature.
CaseAmbient TemperatureMaximum Temperature
(Hydrogen, °C)
Maximum Temperature
(Tank Shell, °C)
Maximum Temperature
(Outer Surface, °C)
7−20 °C62.659.850.3
80 °C70.667.858.3
Ref20 °C79.176.467.7
940 °C87.685.077.6
Table 6. Maximum temperature for each region of tube skid by initial temperature.
Table 6. Maximum temperature for each region of tube skid by initial temperature.
CaseInitial TemperatureMaximum Temperature
(Hydrogen, °C)
Maximum Temperature
(Tank Shell, °C)
Maximum Temperature
(Outer Surface, °C)
10−20 °C55.653.447.4
110 °C66.964.457.1
Ref20 °C79.176.467.7
1240 °C91.888.979.1
Table 7. Maximum temperature for each region of tube skid by initial pressure.
Table 7. Maximum temperature for each region of tube skid by initial pressure.
CaseInitial PressureMaximum Temperature
(Hydrogen, °C)
Maximum Temperature
(Tank Shell, °C)
Maximum Temperature
(Outer Surface, °C)
135 bar81.578.669.5
1410 bar80.177.268.4
Ref15 bar79.176.467.7
1520 bar78.175.366.7
Table 8. Range of parameters for factorial analysis.
Table 8. Range of parameters for factorial analysis.
ParameterLow Limit ValueUpper Limit Value
Charging Time60240
Inlet Temperature−2040
Ambient Temperature−2040
Initial Temperature−2040
Initial Pressure520
Table 9. Factorial analysis for hydrogen tank filling conditions.
Table 9. Factorial analysis for hydrogen tank filling conditions.
CaseCharging TimeInlet TemperatureAmbient TemperatureInitial TemperatureInitial Pressure
160−20−20−2020
2240−20−20−205
36040−20−205
424040−20−2020
560−2040−205
6240−2040−2020
7604040−2020
82404040−205
960−20−20405
10240−20−204020
116040−204020
1224040−20405
1360−20404020
14240−2040405
15604040405
1624040404020
Table 10. Factorial analysis for maximum temperature of inner gas.
Table 10. Factorial analysis for maximum temperature of inner gas.
SourceDFAdj SSAdj MSF-Valuep-Value
  Model510,368207411.960.001
  Linear510,368207411.960.001
Charging Time13978397822.940.001
Inlet Temperature11131130.650.438
Ambient Temperature12382238213.740.004
Initial Temperature13888388822.430.001
Initial Pressure1770.040.846
Error101734173
Total1512,102
Table 11. Factorial analysis for maximum temperature of tank shell.
Table 11. Factorial analysis for maximum temperature of tank shell.
SourceDFAdj SSAdj MSF-Valuep-Value
  Model510,486209714.320.000
  Linear510,486209714.320.000
Charging Time13621362124.730.001
Inlet Temperature163630.430.526
Ambient Temperature12736273618.680.002
Initial Temperature14042404227.610.000
Initial Pressure124240.160.695
Error101464146
Total1511,951
Table 12. Factorial analysis for maximum temperature of outer surface.
Table 12. Factorial analysis for maximum temperature of outer surface.
SourceDFAdj SSAdj MSF-Valuep-Value
  Model59222184415.610.000
  Linear59222184415.610.000
Charging Time12077207717.580.002
Inlet Temperature11181181.0000.341
Ambient Temperature13539353929.960.000
Initial Temperature13488348829.520.000
Initial Pressure1000.0000.985
Error101181118
Total1510,403
Table 13. Response surface analysis for hydrogen tank filling conditions.
Table 13. Response surface analysis for hydrogen tank filling conditions.
CaseCharging Time (min)Ambient Temperature (°C)Initial Temperature (°C)
160−20−20
2240−20−20
36040−20
424040−20
560−2040
6240−2040
7604040
82404040
9601010
102401010
11150−2010
121504010
1315010−20
141501040
151501010
Table 14. Response surface analysis for maximum temperature of inner gas.
Table 14. Response surface analysis for maximum temperature of inner gas.
SourceDFAdj SSAdj MSF-Valuep-Value
  Model68205136813700.000
  Linear37246241524210.000
Charging Time (tchr)12649264926550.000
Ambient Temperature (Tamb)11808180818120.000
Initial Temperature (Tini)12789278927950.000
Square11711711710.000
tchr × tchr11711711710.000
Interaction27883943950.000
tchr × Tamb13653653660.000
tchr × Tini14234234240.000
Error13131
Total198218
Table 15. Response surface analysis for maximum temperature of tank shell.
Table 15. Response surface analysis for maximum temperature of tank shell.
SourceDFAdj SSAdj MSF-Valuep-Value
  Model67650127515490.000
  Linear36785226227480.000
Charging Time (tchr)12233223327130.000
Ambient Temperature (Tamb)11866186622670.000
Initial Temperature (Tini)12686268632630.000
Square11371371670.000
tchr × tchr11371371670.000
Interaction27273644420.000
tchr × Tamb13403404120.000
tchr × Tini13883884710.000
Error13111
Total197661
Table 16. Response surface analysis for maximum temperature of outer surface.
Table 16. Response surface analysis for maximum temperature of outer surface.
SourceDFAdj SSAdj MSF-Valuep-Value
  Model7624289223670.000
  Linear35668188950160.000
Charging Time (tchr)11213121332200.000
Ambient Temperature (Tamb)12244224459570.000
Initial Temperature (Tini)12212221258720.000
Square278391040.000
tchr × tchr13232850.000
Tini × Tini144110.000
Interaction24952486570.000
tchr × Tamb12292296070.000
tchr × Tini11672677080.000
Error1250
Total196246
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

Lee, Y.; Hur, D.-J.; Noh, J.-H. Numerical Analysis of Temperature Distribution During Charging Process of Vertically Installed Hydrogen Tanks. Appl. Sci. 2025, 15, 1193. https://doi.org/10.3390/app15031193

AMA Style

Lee Y, Hur D-J, Noh J-H. Numerical Analysis of Temperature Distribution During Charging Process of Vertically Installed Hydrogen Tanks. Applied Sciences. 2025; 15(3):1193. https://doi.org/10.3390/app15031193

Chicago/Turabian Style

Lee, Yeseung, Deog-Jae Hur, and Jung-Hun Noh. 2025. "Numerical Analysis of Temperature Distribution During Charging Process of Vertically Installed Hydrogen Tanks" Applied Sciences 15, no. 3: 1193. https://doi.org/10.3390/app15031193

APA Style

Lee, Y., Hur, D.-J., & Noh, J.-H. (2025). Numerical Analysis of Temperature Distribution During Charging Process of Vertically Installed Hydrogen Tanks. Applied Sciences, 15(3), 1193. https://doi.org/10.3390/app15031193

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