Next Article in Journal
Investment Analysis of Low-Carbon Yard Cranes: Integrating Monte Carlo Simulation and Jump Diffusion Processes with a Hybrid American–European Real Options Approach
Next Article in Special Issue
Design and Simulation of an Integrated Process for the Co-Production of Power, Hydrogen, and DME by Using an Electrolyzer’s System
Previous Article in Journal
Robust Control of a Zeta Converter: A Feedback Linearization Approach with Digital PWM Implementation
Previous Article in Special Issue
Performance and Emissions of Spark-Ignition Internal Combustion Engine Operating with Bioethanol–Gasoline Blends at High Altitudes Under Low- and High-Speed Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hydrogen Properties and Their Safety Implications for Experimental Testing of Wing Structure-Integrated Hydrogen Tanks

by
Javed A. Butt
* and
Johannes F. C. Markmiller
Chair of Aircraft Engineering, Institute of Aerospace Engineering, TUD Dresden University of Technology, 01062 Dresden, Germany
*
Author to whom correspondence should be addressed.
Energies 2025, 18(8), 1930; https://doi.org/10.3390/en18081930
Submission received: 24 February 2025 / Revised: 1 April 2025 / Accepted: 8 April 2025 / Published: 10 April 2025
(This article belongs to the Special Issue Renewable Fuels: A Key Step Towards Global Sustainability)

Abstract

:
Hydrogen is a promising candidate for addressing environmental challenges in aviation, yet its use in structural validation tests for Wing Structure-Integrated high-pressure Hydrogen Tanks (SWITHs) remains underexplored. To the best of the authors’ knowledge, this study represents the first attempt to assess the feasibility of conducting such tests with hydrogen at aircraft scales. It first introduces hydrogen’s general properties, followed by a detailed exploration of the potential hazards associated with its use, substantiated by experimental and simulation results. Key factors triggering risks, such as ignition and detonation, are identified, and methods to mitigate these risks are presented. While the findings affirm that hydrogen can be used safely in aviation if responsibly managed, they caution against immediate large-scale experimental testing of SWITHs due to current knowledge and technology limitations. To address this, a roadmap with two long-term objectives is outlined as follows: first, enabling structural validation tests at scales equivalent to large aircraft for certification; second, advancing simulation techniques to complement and eventually reduce reliance on costly experiments while ensuring sufficient accuracy for SWITH certification. This roadmap begins with smaller-scale experimental and numerical studies as an initial step.

1. Introduction

The aerospace industry consistently strives for innovations that enhance fuel efficiency [1,2,3], reduce aircraft weight [4,5,6,7,8], and improve overall performance [9,10,11]. A notable advancement in this realm is the concept of Wing Structure-Integrated high-pressure Hydrogen Tanks (SWITHs), pronounced sweets, which are embedded within an aircraft’s wing structure. An illustrative depiction of a SWITH is provided in Figure 1.
The primary distinction between integrated and externally stored fuel tanks lies in their intended purpose. Integrated tanks are engineered to bear external loads and provide support to the aircraft’s load-bearing components, whereas external tanks are employed exclusively for fuel storage and supply. They do not constitute an integral component of the aircraft’s structure and, consequently, are not designed to support the load-bearing structure by absorbing loads. In addition to wing-integrated hydrogen tanks, hydrogen storage can also be incorporated into the fuselage [12,13,14,15,16,17,18,19,20,21,22]. The utilization of hydrogen tanks, whether integrated into the wing or the fuselage, can take the form of liquid or gaseous hydrogen. However, the present contribution is focused on gaseous SWITHs. The main advantage of using hydrogen lies in its potential for more environmentally friendly aviation, as evidenced by numerous studies [23,24,25,26,27]. Moreover, hydrogen’s significance is reinforced by its diverse applications across multiple fields. In the mobility sector alone [28,29,30], it has the potential to power various vehicles, including ships [31,32], motorcycles [31,33,34,35,36,37,38], cars [39,40,41,42,43], heavy-duty trucks [33], submarines [32,44,45,46], trains [31,33,47], and unmanned aerial vehicles (UAVs) [33,48,49,50,51,52].
Figure 1. Illustrative example of a SWITH, similar to APUS i-2 [53].
Figure 1. Illustrative example of a SWITH, similar to APUS i-2 [53].
Energies 18 01930 g001
In the context of aviation, integrating hydrogen storage directly into the aircraft structure offers additional benefits beyond propulsion. The key benefit of integrating tanks as a part of the aircraft’s integral design is the liberation of space that would otherwise be allocated for fuel storage. This freed-up space can then be repurposed for other uses, such as installing additional seating to increase passenger capacity. Alternatively, it could carry more cargo or large, cumbersome items that do not exceed the size of the original tanks. The notion of SWITHs has evolved from a mere scientific concept to a tangible reality. For instance, APUS [53], a company specializing in aircraft manufacturing, is currently engaged in the development of its i-2 model, with the objective of introducing the first commercially available SWITH [54]. Notwithstanding these promising advancements, the relatively nascent SWITH concept remains confronted with challenges pertaining to manufacturing, safety, ambiguous regulatory frameworks, and a paucity of extant literature.
Given the innovative nature of SWITHs, scholarly investigation—particularly regarding hydrogen’s role as a filling agent in structural validation tests for aircraft-scale SWITHs—remains scarce. This gap in academic discourse highlights the importance of this study, which aims to assess the implications of using hydrogen in such tests and its impact on their feasibility. Furthermore, this understanding is especially beneficial when designing new SWITHs, as it is pivotal for addressing key hydrogen safety-related challenges. To achieve this, the study first provides an overview of hydrogen’s fundamental properties, with an emphasis on its beneficial characteristics. This is followed by an analysis of its safety-relevant aspects, supported by numerical simulations and experimental findings. Based on this foundation, the feasibility of conducting static experimental structural validation tests on aircraft-scale SWITHs, with hydrogen as the filling agent, is evaluated. If such testing is found to be impractical in its current form, potential pathways toward its realization are proposed.
Given the complexity and innovative nature of SWITHs, a collaborative effort among researchers is essential to address the challenges associated with their development. As of the writing of this paper, no formal standards or regulations exist for the certification of SWITHs [55]. Consequently, the findings from this study may serve as a valuable reference for the future development of such standards and regulations. The necessity of well-defined standards becomes particularly evident in the context of experimental tests and simulations on SWITHs, as they establish critical benchmarks for validation, safety, and reliability. Without such guidelines, ensuring consistency in testing methodologies and assessing structural integrity under real-world conditions becomes significantly more challenging. Establishing and adhering to structured standards allows engineers and specialists to conduct evaluations in a systematic, transparent, and safety-oriented manner [56,57,58,59,60,61].

2. Fundamental Properties of Hydrogen: An Overview

To understand hydrogen’s implications for experimental structural validation tests of aircraft-sized SWITHs, its basic properties are first introduced. This understanding is crucial in recognizing why hydrogen’s unique characteristics make it a promising alternative fuel. Safety-related properties are examined separately in Section 3. For convenience, all key information from this section is compiled in Table 1.
Table 1. Fundamental properties of hydrogen for aerospace and energy applications.
Table 1. Fundamental properties of hydrogen for aerospace and energy applications.
PropertyDescription
Atomic structure
  • Hydrogen (H) is the smallest known atom, with a diameter of approximately 0.07 nm [62,63].
Molecular form
  • It exists primarily as a H2 molecule [62,63].
  • The H atom itself is highly reactive [64,65,66].
  • Generally, references to hydrogen denote the H2 molecule rather than the H atom.
Chemical properties
  • The H atom forms a stable bond with another H atom to create H2. This stable bond results from fully occupied valence electrons, leading to the low reactivity of H2 [67].
  • The H atom has a relatively low electronegativity of 2.2 [62,63], which describes an atom’s ability to attract electrons when forming chemical bonds.
Sensory properties
Safety properties
  • Non-toxic [69,71,72].
  • Non-carcinogenic [62,63].
  • Not radioactive [63].
  • Not caustic [62,63].
Production
  • Renewable production is possible through water electrolysis powered by wind, solar, or other renewable energy sources [73,74,75,76].
Thermodynamic properties
  • Hydrogen exhibits a negative Joule–Thomson coefficient at high pressures [77,78,79]. This implies that, under constant enthalpy h, an increase in pressure results in a decrease in temperature (cooling during relaxation in a throttle [62]), and vice versa. This relationship can be mathematically expressed as follows:
    μ JT = T p h
    where T represents temperature, p denotes pressure, h indicates the constant enthalpy, and denotes the derivative or change in temperature with respect to pressure. Hydrogen also deviates from ideal gas behavior.
Physical properties
  • Hydrogen is the lowest-density element (it is approximately 14 times lighter than air) [62,63].
  • It primarily exists in liquid and gaseous states.
  • At 0 °C and 1.01325 bar, hydrogen has a high speed of sound ( a H 2 = 1261.1 m / s [62]) compared to air ( a air = 331.5 m / s [80]).
  • Extensive research is ongoing into alternative storage forms [81,82,83], including physical and chemical adsorption (metallic hydride) [84,85,86], supercritical or cryo-compressed storage [87,88,89], and other methods, as illustrated in Figure 2.
Abundance
  • Hydrogen is plentiful, including in Earth’s atmosphere [69,70,72,90,91,92].
  • It is a component of many fossil fuels, including methane (CH4), methanol (CH3OH), ethene (C2H4), propane (C3H8), and benzene (C6H6) [62].
Regulations
  • Safe handling and storage standards are available and are comparable to those for natural gas [62,63,73,93,94].
While Table 1 provides essential information about hydrogen, one key aspect warrants closer examination: its gravimetric energy density. Gravimetric energy density defines the usable energy per unit mass, with higher values enabling greater energy storage at lower weight. Due to the ongoing environmental crisis [95,96], most mobility sectors are eager to find an energy medium that offers both a high energy content for economic reasons and a low weight for environmental reasons. The lighter the energy carrier, the less mass needs to be transported, thereby enabling lower overall power consumption.
Figure 2. Illustrative example of different methods for storing hydrogen [81].
Figure 2. Illustrative example of different methods for storing hydrogen [81].
Energies 18 01930 g002
In its pure form, hydrogen possesses a high gravimetric energy density [81,97,98,99]. However, the gravimetric energy densities achieved in practical applications are significantly lower. This reduction is primarily due to storage system specifications. For practical applications, hydrogen must be stored within a tank. In the case of compressed gaseous hydrogen (CGH2), pressure is a major factor influencing the gravimetric energy density [100,101]. However, there are several considerations regarding pressure and tank design [102,103,104,105,106,107]. Firstly, there is a technical limit on how much pressure can be contained within a tank. Secondly, higher operating pressure exerts more stress on the tank, dictating appropriate tank materials [44,108,109,110]. Furthermore, the mass necessary for pressure containment reduces the overall gravimetric energy density of the hydrogen storage system, as the tank’s mass does not contribute to the energy content. Consequently, if the tank is very heavy but does not allow for the storage of much hydrogen mass, the resulting gravimetric energy density is low. A comparison of gravimetric energy densities between pure forms and practical storage systems for commonly used fuels is presented in Figure 3. In this figure, the suffix G denotes a gaseous state, L denotes a liquid state, and NG denotes natural gas.
When considering only the gravimetric energy density of pure hydrogen, Figure 3 clearly shows that hydrogen provides more energy per unit mass than other commonly used fuel sources, such as natural gas, gasoline, and diesel.
In pursuit of a more nuanced analysis regarding the non-ideal gas behavior of hydrogen, the real gas factor [111], also known as the compressibility factor, is introduced. It indicates the degree to which a gas deviates from ideal gas behavior. It can be represented mathematically by Equation (2), as shown below, where the variables Z and m signify the real gas factor and the mass of the specific gas, respectively.
Z = m ideal m real
When Z equals one ( Z = 1 ), there is no disparity between real and ideal gas behaviors. If Z exceeds one ( m ideal > m real Z > 1 ), the ideal gas equation overestimates the mass, whereas, if Z is less than one ( m ideal < m real Z < 1 ), the equation underestimates the mass. Real gas factors are typically determined empirically and provided as regression-derived analytical equations or terms. An example for hydrogen is outlined in [112] using Equation (3), with the accompanying constants shown in Table 2. The variables p , T , ρ , and R represent pressure, temperature, density, and the ideal gas constant, respectively, while a i , b i , and c i are regression model coefficients. Equation (3) extends the ideal gas equation [113] by incorporating regression terms that account for real gas behavior, which varies with both pressure and temperature.
Z ( p , T ) = p ρ R T = 1 + i = 1 9 a i 100 K T b i p 1 MPa c i
Figure 4 illustrates the deviation of hydrogen from ideal gas behavior at various temperatures, with compression from 0 to 700 bar , represented by real gas factors [62]. The discrepancy between ideal and real gas behavior increases predominantly with rising pressure. Additionally, the effect of temperature between 50 K and 100 K follows a distinctly non-linear pattern.
A custom tool was developed to analyze the real gas factor and density across a specified range of pressure and temperature. This is particularly useful when incorporating fuel mass into simulation models, such as finite element (FE) models [117,118,119]. The tool features an interactive three-dimensional (3D) visualization, allowing for a detailed examination of temperature–pressure interactions and their impact on hydrogen density. A depiction of the tool in its eye-friendly dark mode, intended to reduce visual strain and fatigue [120,121], is shown in Figure 5. The interactive version is freely accessible via GitHub (https://jav-ed.github.io/H2_Plot/, accessed on 10 March 2025). Additionally, the source code is available on GitHub (https://github.com/jav-ed/H2_Plot, accessed on 10 March 2025), allowing users to customize parameters such as pressure and temperature resolution.
To summarize, the essential properties of hydrogen are cataloged in Table 1, while safety aspects will be examined in subsequent sections. Preliminary assessments indicate that hydrogen possesses primarily favorable properties. However, its invisibility to the human eye due to its small molecular size, along with potential measurement difficulties, remains a challenge [122,123,124]. Additionally, an established analytical function was introduced to facilitate the calculation of hydrogen mass across various pressure and temperature ranges.

3. Safety-Critical Characteristics of Hydrogen

One key argument for why a thorough understanding of hydrogen properties is essential for structural validation tests is the role that standards play in ensuring safety and reliability [56,57,58,59,60,61]. In commercial aircraft, standards are fundamental for maintaining the consistency, reliability, and safety of structural validation procedures. Similarly, for SWITHs, well-defined standards would provide crucial guidance on the conditions under which hydrogen must be stored, handled, and tested to accurately replicate operational scenarios. Moreover, if hydrogen poses specific hazards, standards would ensure that these risks are properly identified, assessed, and mitigated. They would outline important factors contributing to these risks and, where available, methods to reduce them. However, since no specific standards for SWITHs currently exist, such predefined guidance is unavailable [55].
Given that SWITHs encompass both aerospace and high-pressure vessel domains, existing standards from these fields can offer valuable insights [125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142]. They may help identify relevant structural validation tests, highlight the unique properties of hydrogen, and provide general guidance on conducting such tests. One such valuable insight can be gained by examining ISO 11119-3:2020 [143], which specifies the tests required for high-pressure cylinder certification. The mandatory and optional structural tests specified in ISO 11119-3:2020 [143] offer an informative overview of additional tests that SWITHs may be subject to, beyond the standard requirements for conventional aircraft structure. Beyond the types of tests and the corresponding hydrogen parameters, ISO 11119-3 states the following in paragraph 8.5.1.1: When carrying out the pressure test, a suitable fluid shall be used as the test medium. This can include liquids such as water or oil and gases such as air or nitrogen. This stipulation allows for various filling media in pressure approval tests. In the context of SWITHs, the filling medium refers to the substance stored and pressurized within the wing tanks. Additionally, the standard permits variations in testing media for specific tests. For instance, natural gas is allowed in the permeability test (5.5.12.2), whereas the pneumatic cycle test (5.5.16.1) specifically requires hydrogen to be the filling medium. These findings indicate that only certain experimental tests allow for alternative filling media.
Beyond consulting standards, the need for a thorough understanding of the filling medium for experimental tests becomes apparent when considering the factors listed in Table 3. These factors represent some key aspects that should be considered when selecting a filling medium, as they may influence structural properties, human health, measurement accuracy, test bench safety, availability, environmental impact, and economic efficiency.
While the factors outlined in Table 3 underscore the necessity for a comprehensive understanding of the medium used in SWITHs, some of them have already been addressed in Section 2. Based on the findings so far, it can be inferred that only tests requiring an alternative gas would expose potential hydrogen-related hazards. From this perspective, in each potentially hazardous test, hydrogen could be replaced with an alternative filling agent. Following this reasoning, one might initially conclude that all tests requiring hydrogen would be inherently safe and pose no risk to humans or the environment. However, this assumption requires further scrutiny, particularly in the context of aircraft-sized SWITHs. Resolving this uncertainty is one of the main objectives of this work. To clarify the potential risks and challenges associated with hydrogen use, it is essential to first establish some fundamental knowledge and definitions. Sigloch [80] provides the following key concepts:
  • Deflagration: a combustion process where the flame front propagates at a speed less than that of sound.
  • Explosion: the rapid, nearly instantaneous combustion of combustible materials or explosives, resulting in a large volume of combustion gas that violently displaces the surrounding air.
  • Detonation: A combustion process occurring at supersonic speeds, accompanied by a strong pressure increase and anticipated pressure shocks. Due to the intensity of this process, hearing protection is mandatory in proximity to such events.
The combustion speeds and associated pressure increases that are to be expected for the different combustion processes are listed in Table 4. The magnitude of the destructive power for the given pressure increases can be estimated using Table 5.
A vital consideration in experimental investigations involving hydrogen is its potential for combustion, which can manifest as deflagration, explosion, or detonation. Notably, hydrogen alone is not inherently flammable [62,144,145,146,147]. However, under specific conditions, hydrogen can become flammable or even detonate. The crucial factors, as outlined in Table 6, create the conditions necessary for hydrogen’s combustibility.
The minimum ignition energy varies with the air-to-hydrogen ratio. However, even a small spark or static discharge can ignite hydrogen in certain gas mixtures. For instance, a static discharge of 20–30 mJ from rubbing against a carpet and touching a doorknob is enough to ignite a hydrogen–air mixture [62]. Figure 6 illustrates the minimum ignition energy (mJ) of hydrogen (H2) in humid air (90% relative humidity) and dry air (0% relative humidity). This confirms that a static discharge as low as 20 mJ can ignite a hydrogen–air mixture. For comparison with other fuels, such as methane (CH4) and propane (C3H8), the literature provides similar plots of minimum ignition energy across various air compositions [62,63,73]. A comparison of hydrogen, methane, and propane indicates that hydrogen has a lower minimum ignition energy and a broader flammability range.
For further important considerations, let us define the following two technical terms:
  • Flash point: For a flammable liquid, this refers to the lowest temperature at which, under specified conditions, sufficient vapors are produced to form an ignitable vapor–air mixture above the liquid surface upon external ignition. Notably, if the ignition source is removed, the flames extinguish [62,63].
  • Ignition temperature: this denotes the lowest temperature at which spontaneous ignition of the fuel occurs in an open vessel [62,63].
In both cases, ignition can only occur if a certain amount of air is present. Table 7 offers valuable insights into the ignition or explosion behavior of hydrogen and other fuels. Compared to other gases, hydrogen’s ignition temperature is relatively high, yet the flammability range of a hydrogen–air mixture is exceptionally wide [151,152,158]. While this broad range may be manageable with various safety measures, the fact that hydrogen has the lowest minimum ignition energy among the fluids listed in Table 7 poses a significant challenge for preventing both unintentional and accidental ignitions [151,158,159,160,161,162,163].
Further complicating safety considerations is the fact that hydrogen is invisible to the naked eye. However, experimental investigations conducted by [122,123,124,164] have successfully recorded hydrogen flames. A visual representation of these flames is shown in Figure 7, offering insights into the observable features of hydrogen combustion. This image was captured at a laminar jet velocity of 47 m / s and a Reynolds number of 837 using a Sony DSC D700 (Digital Still Camera, 1344 × 1024 pixels), an f/2.4 aperture, and no filter. Contrary to the common misconception that hydrogen flames are invisible, Figure 7 demonstrates that they are, in fact, visible. Notably, while hydrogen flames are indeed visible, they exhibit lower luminosity than those of burning hydrocarbons [164]. This difference in visibility has significant implications for hydrogen safety and detection methods [122,123,124]. The visibility of hydrogen flames is further influenced by the oxygen content in the mixture, as illustrated in Figure 8. The link between flame visibility and mixture composition provides valuable insights for practical safety measures.
When examining hydrogen flame visibility, the equivalence ratio, denoted as ϕ , plays a crucial role. ϕ is defined as the ratio of the actual air–fuel ratio to the stoichiometric air–fuel ratio for combustion [165]. An equivalence ratio of ϕ = 1 corresponds to stoichiometric combustion, while lower values of ϕ indicate suboptimal combustion conditions. The experimental results show a clear correlation between ϕ and flame visibility: as ϕ increases, the flame becomes more visible. The experimental conditions for these observations are detailed in Figure 8. As [164] highlights, low ambient lighting enhances hydrogen flame visibility, a critical factor in practical flame detection.
Table 8 contains detailed data on the limits of hydrogen–air mixtures for various types of ignition and other relevant parameters. Under normal conditions (0 °C, 1.01325 bar), hydrogen is flammable at concentrations between 4 and 75.6 % and detonable between 18 and 58.9 % [63].
Beyond the challenges related to air content outlined in Table 8, Table 6 highlights two exceptional cases of hydrogen reactivity. First, hydrogen and chlorine can react solely through light irradiation, producing a loud, explosive bang known as the chlorine oxyhydrogen reaction [62,63]. Second, hydrogen undergoes a similarly explosive reaction with fluorine [62,63]. Further studies on the explosion limits of various hydrogen–oxygen mixtures can be found in the literature, such as [166,167,168,169,170,171,172]. These cases underscore the diverse and potentially hazardous reactive properties of hydrogen. The reactive nature of hydrogen not only manifests in gaseous interactions but also in its interactions with solid materials.
Therefore, an important consideration in hydrogen-related SWITH simulations and physical tests is the material compatibility of hydrogen. Hydrogen can accumulate in metal lattices, leading to embrittlement, which degrades structural properties [44,173,174,175,176,177]. Hydrogen incorporation into metal lattices occurs when hydrogen molecules dissociate into atoms at the metal’s surface [63,81]. Certain pure metals, including palladium, magnesium, lanthanum, and aluminum, as well as specific alloys such as TiNi-Ti2Ni and Mg-Mg2Ni, can store hydrogen [63]. Hydrogen storage in metal lattices can also be used to produce highly pure hydrogen (99.99%) [81,178,179]. Hence, metal lattices can act as filters, turning a potential challenge into a practical benefit for hydrogen purification.
Although plastics are not susceptible to hydrogen embrittlement [102,103,104,105,106,107], they present significant challenges in retaining hydrogen. In the worst-case scenario, a filled tank could rapidly deplete due to hydrogen permeation through the vessel walls. If a leaking hydrogen tank is stored in an enclosed space, such as a garage, the accumulation of hydrogen could create an undetected explosive atmosphere, where even a small spark or open flame could trigger a fire or explosion, posing a serious risk to occupants. Therefore, two of the most critical properties to consider for hydrogen storage are its flammability and diffusion behavior. For a more detailed discussion on hydrogen diffusion in polymers, see [180].
In summary, hydrogen is not flammable without the addition of oxygen. However, a wide range of hydrogen–air mixtures can ignite. Furthermore, even a very small static discharge is sufficient to ignite a flammable gas mixture, leading to a potential detonation. Hydrogen affects structural properties, particularly in metals, by causing embrittlement; while plastics are not susceptible to embrittlement, they face significant challenges in retaining hydrogen. Given hydrogen’s diffusion behavior and flammability, even a minor undetected leak in a highly pressurized tank could endanger both personnel and instruments during experimental tests for SWITH certification. Consequently, the following section naturally transitions to the experimental and numerical validation of these phenomena.

4. Safety-Focused Hydrogen Experiments and Simulations for SWITHs

This section bridges the theoretical understanding of hydrogen with practical insights from experimental and simulation investigations. It evaluates the feasibility of conducting structural validation tests with hydrogen, which may be essential for future SWITH certification. A critical phenomenon in this context is the formation of shock waves, which arise when local flow velocities exceed the speed of sound, as described in [80,181] and observed experimentally in [182,183]. Shock waves are complex phenomena [184] and their description will be limited to what is necessary for understanding the presented results. Shock waves primarily affect their surroundings in three distinct ways. They generate extremely high pressures, while temperatures can reach extreme levels, sometimes exceeding several thousand Kelvin (e.g., 3000 K [185] and higher [186]). Additionally, rapid and significant pressure changes create airborne noise, necessitating the use of hearing protection.
Xu et al. [185] simulated these shock wave effects by compressing hydrogen to 250 bar in a pressure vessel and releasing it into the free atmosphere through a nozzle. The simulation output, shown in Figure 9, depicts the system at various time intervals after the hydrogen exits the nozzle. To analyze this output, the Mach number ( M a ) is defined using Equation (4) below, where a and v represent the speed of sound and local velocity, respectively. The Mach number expresses the ratio of local velocity to the speed of sound. For example, M a = 5 indicates that the local velocity is five times the speed of sound.
M a = v a
As shown in Figure 9, the Mach scale in this simulation extends up to seven, reflecting the range of the computed results. Notably, shock waves—and their associated consequences—begin to occur at Mach 1. As the Mach number increases, these effects, including high temperatures, high pressures, and increased noise, intensify significantly. For a more in-depth interpretation of these results, Table 9 can be consulted.
Each medium has a characteristic speed of sound that determines the propagation of pressure changes and sound waves through it. This relationship is expressed by Equation (5), where a, p, and ρ denote the speed of sound, pressure, and density, respectively [80], as follows:
a = d p d ρ
A higher speed of sound in a medium corresponds to faster information transmission. For instance, when one end of a steel rod is struck, the sound propagates to the other end at 5170 m / s , making it audible almost instantly. The speed of sound in hydrogen at 20 °C and 1 bar is 1300 m / s , which is significantly higher than in air ( 343 m / s ). Consequently, information exchange in hydrogen occurs more rapidly than in air. This characteristic is crucial for understanding the behavior of pressurized hydrogen released into the atmosphere. Air undergoes a shock wave at 343 m / s , causing a significant temperature rise. In contrast, hydrogen can reach nearly four times this speed before experiencing a shock wave. A simplified explanation for the pronounced temperature increase during shock waves is as follows: when air molecules cannot adapt quickly to pressure changes because the speed at which pressure changes propagate is limited, they are forced to change abruptly. This phenomenon is analogous to a supersonic aircraft encountering air molecules before they can adjust to its approach. The resulting rapid compression and frequent, intense collisions convert substantial kinetic energy into thermal energy, yielding a marked temperature rise. This process predominantly occurs in compressible media like gases, explaining the prominence of shock waves in air and hydrogen. These temperature effects are visualized in Figure 10, which presents the temperature contours associated with Figure 9. Temperatures exceeding 1000 K are typically associated with explosions [186,187,188,189].
The provided explanation, when applied to the described outcomes, indicates that, as hydrogen flows into the free environment, its outflow velocity already exceeds the speed of sound in the surrounding air. Consequently, the air experiences a shock wave, leading to rapid compression and a significant temperature increase. Hydrogen itself, however, would not initially exceed its own speed of sound and, therefore, would not generate a shock wave upon release. The elevated temperatures near the hydrogen, induced by the air shock wave, may not immediately trigger detonation. Subsequently, as hydrogen disperses into the air and diffusion occurs, an ignitable concentration of the oxygen–hydrogen mixture is eventually reached [190]. At this point, the surrounding high temperature becomes sufficient to cause ignition, which, under certain conditions, can escalate to detonation. This explanation is a simplified model illustrating core concepts. The actual physics of shock wave formation and temperature rise involve complex thermodynamics. For a more detailed discussion, refer to advanced texts on nonlinear compressible fluid dynamics [191].
Ignitions that occur without external ignition sources are referred to as self-ignitions or spontaneous ignitions [192]. Various theoretical models for self-ignition have been proposed, including the reverse Joule–Thomson effect, electrostatic ignition, brush and corona discharges, diffusion ignition, sudden adiabatic compression, hot surface ignition, mechanical friction, and impact ignition [151,193,194,195]. Despite the variety of proposed mechanisms, not all are equally probable in real-world scenarios. Compression ignition, Joule–Thomson expansion, diffusion ignition, and ignition by hot surfaces are generally considered unlikely for most unintentional hydrogen releases at ambient temperatures [193]. However, it is important to recognize that several of these mechanisms may collectively contribute to a self-ignition event [193]. Among these mechanisms, diffusion ignition, as proposed by Wolański and Wójcicki [196], has received significant attention. This model describes a spontaneous ignition phenomenon that occurs when pressurized hydrogen is released into a chamber containing either pure oxygen or air [195]. Interestingly, ref. [193] concluded that diffusion ignition, along with compression ignition, Joule–Thomson expansion, and ignition by hot surfaces, is unlikely for most unintentional hydrogen releases at ambient temperature.
Nonetheless, despite this assessment, extensive research has been conducted on the diffusion ignition mechanism. One notable example of intensive experimental investigations into the outflow of pressurized hydrogen into the free environment was conducted by [93]. Additionally, numerous studies using cylindrical shock tubes have been carried out by various researchers [197,198,199,200,201,202,203,204,205,206,207]. Comprehensive reviews and summaries of spontaneous ignition mechanisms in pressurized hydrogen released through tubes are available in [151,195]. The findings from these studies are not only valuable for conducting experimental investigations with hydrogen but also have major implications for the design of SWITHs. In the following paragraphs, these findings are presented in detail, accompanied by an explanation of their relevance to SWITH design and experimental procedures.
When pressurized hydrogen is released from its vessel, certain geometrical aspects have a significant impact on the subsequent spontaneous ignition. One crucial factor is the tube length. Both numerical simulations and experimental studies on releasing pressurized hydrogen indicate that the longer the pipe, the greater the likelihood of spontaneous ignition occurrence is [151,194,195,208]. However, when the tube length exceeds a certain threshold, the likelihood of self-ignition in high-pressure hydrogen decreases [194]. According to [207], this reduction occurs only when the tube length exceeds 1700 mm ; while longer tubes are economically preferable [98,209,210], their potential to increase the risk of self-ignition complicates the determination of an optimal length. Existing experimental studies on self-ignition focus on shorter tubes. However, for SWITH design, it may be necessary to investigate significantly longer configurations. The need for tubes ranging from approximately 5 m to 80 m [211] is driven by economic and practical considerations. Maximizing flight range is a key economic goal, while the length of wing-integrated tanks in commercial aircraft significantly exceeds that of tubes typically used in experimental research.
The preference for shorter tubes (less than 5 m ) in current experiments is primarily due to cost and time constraints faced by researchers. Previous reviews [151,194,195,207] have indicated that experimental tests are generally conducted with tube lengths of ≤ 4200 mm . Notably, tube lengths of ≤ 1000 mm are more common than the comparatively long 4200 mm tubes used in [202]. A comprehensive list of experimental parameters, including the tube lengths used in various studies, is provided in Table 10. This table highlights the limitations of current experimental setups and underscores the need for tests that more closely reflect the actual dimensions of aircraft-sized SWITHs.
Depending on the outcome of self-ignition experiments assessing aircraft-relevant tube lengths and economic feasibility, numerical optimization [212,213,214,215,216,217,218,219,220,221,222] may be required to determine optimal tube dimensions. The need for optimization is further highlighted by the influence of multiple geometrical parameters, beyond just length, on self-ignition behavior. In addition to tube length, diameter is another critical factor. Studies have shown that tubes with smaller diameters are more prone to spontaneous ignition [190,194,195]. Consequently, based on our current understanding, a larger diameter is generally recommended to mitigate the risk of self-ignition. The cross-sectional area is another important geometrical parameter influencing spontaneous ignition [223,224]. Variations in the local cross-section, whether decreases or increases, lower the critical release pressure threshold for ignition compared to tubes with constant cross-sections [195,223,224]. The overall tube shape has little impact on the minimum burst pressure needed for spontaneous ignition [195]. However, in non-circular cross-section tubes, turbulent flow in the corners can enhance the mixing of hydrogen and air. This enhanced mixing increases the amount of hydrogen–air mixture formed [225].
Both diameter and cross-section are key parameters in designing SWITH tubes. Along with tube length, these factors determine the total volume of SWITH tubes. For clarity, Table 11 provides a concise summary of these findings.
Hydrogen self-ignition has been observed both in theoretical models and real-world applications [226]. This real-world occurrence highlights the need for strict hydrogen safety measures, especially for experimental validation tests of SWITHs. Consequently, further investigations are imperative to establish and refine safe working protocols. To better understand the recommendation in Table 11 for maintaining low pressure inside hydrogen containers, the study reported in [227] is examined in detail. This study investigated the effects of various oxidizing agents on hydrogen’s spontaneous ignition. The experimental setup involved releasing hydrogen from a high-pressure vessel through a nozzle, similar to previously mentioned studies. However, a key distinguishing feature of this experiment was the controlled environment surrounding the hydrogen jet. Unlike earlier investigations conducted in an open atmosphere, the experiment in [227] introduced a constant flow of oxidizing agents perpendicular to the hydrogen jet. The tested oxidizing agents included air, pure oxygen (O2), nitrous oxide (N2O), and acetylene (C2H2). The experimental apparatus utilized a straight expansion tube with specific dimensions: a diameter of 4 mm and a length of 10 cm . To capture spontaneous ignition, the researchers employed a high-speed camera. Additionally, they measured external overpressures to quantify the effects of the ignition. The experimental findings displayed in Figure 11 illustrate how hydrogen container pressure and ambient gas composition influence the likelihood of self-ignition.
The results depicted in Figure 11 reveal several key insights. First, there is a clear correlation between increasing hydrogen container pressure and a higher probability of self-ignition. At low pressures, self-ignition was avoided in all tested gas environments. This observation is particularly significant, as preventing unintended hydrogen ignition is a key safety requirement for practical applications. When considering only the prevention of self-ignition, the tested ambient gases can be ranked by safety. Oxygen (O2) presents the most hazardous conditions for hydrogen applications; while pure air and nitrous oxide (N2O) are less hazardous than oxygen, they are still considered unsafe for hydrogen applications. Among the tested gases, only acetylene (C2H2), which is itself combustible, met safety requirements for pressures below 230 bar .
Revisiting key points from Section 3 and this section confirms that hydrogen possesses both safety-critical and beneficial properties. Despite potential hazards, hydrogen remains a viable option for practical applications when handled responsibly and with a comprehensive understanding of its unique characteristics [62,73,94,227,228]. Notably, Landucci et al. [228] concluded that the risks associated with compressed hydrogen are comparable to those of liquefied petroleum gas and natural gas. The referenced studies are largely based on specialized testing environments or the automotive industry; while their findings offer valuable insights for SWITHs, they differ significantly from aircraft requirements. Understanding these differences requires a closer look at SWITH-specific constraints. In SWITHs, hydrogen tanks are integrated into the aircraft wings, where they must maintain high internal pressure and volume for optimal energy density. Additionally, they must withstand both internal pressure and external loads. The aviation industry’s pursuit of minimal weight and high safety standards favors the use of mature Type IV tanks [102,104,106,108,109,110,229,230,231,232]. However, this design choice introduces additional safety concerns. In the event of rupture, the high-speed dispersion of composite tank fragments could endanger both personnel and measurement devices. Moreover, composite breakdown may release microscopic particles or fibers into the air. Inhalation of these particles can lead to respiratory problems [233,234], a risk that is particularly acute in confined, poorly ventilated spaces.
The goal of maximizing hydrogen load during flight necessitates optimizing tank pressure and volume. This requirement emphasizes a key distinction between SWITHs and most other applications of compressed hydrogen cylinders: the size of the tanks. For instance, the APUS i-2 [54] has a wingspan of 13.2 m , with approximately ⪆ 50 % of its length occupied by four cylindrical high-pressure tubes. In comparison, the Airbus A350-1000 has a wingspan of about 64 m and a flight range of 16 , 482 km [235]. Assuming 50% of this wingspan could accommodate tubes, the total tube length would be 64 m × 0.5 = 32 m . However, the objectives of maximizing tank size and internal pressure may inadvertently increase the risk of self-ignition, as highlighted earlier in this section.
Another key distinguishing factor for SWITHs, particularly compared to land vehicles, is their operation across a range of altitudes. This variability in operational environment introduces several complex phenomena. At higher altitudes, reduced ambient pressure increases the pressure differential between the hydrogen tanks and the surrounding environment. This greater differential may lead to higher hydrogen release velocities in the event of a leak. Importantly, if the release velocity does not exceed the speed of sound in hydrogen, a hydrogen shock wave is not expected. Additionally, this effect is further complicated by the decrease in the speed of sound in air with altitude. As a result, the probability of an air shock wave forming increases, even at lower hydrogen release velocities. When diffusion causes the released hydrogen to mix with surrounding air at an ignitable concentration, various outcomes—ranging from ignition to detonation—become possible. Conversely, lower air density at higher altitudes [236,237] introduces a counteracting effect. The reduced density may slow or complicate the formation of ignitable concentrations as air and hydrogen diffuse. This adds another layer of complexity to the safety considerations for SWITHs operating at various altitudes.
Furthermore, considering the increased pressure difference and reduced air density in tandem, the behavior of released hydrogen becomes more nuanced. The speed at which hydrogen is released is not only enhanced, due to the decreased ambient pressure, but also due to an increased mean free path for gas particles. In other words, hydrogen molecules can traverse longer distances before colliding with air molecules, leading to a delayed reduction in speed. This delay occurs because fewer collisions reduce the conversion of kinetic energy into thermal energy. Consequently, lower air density at higher altitudes allows hydrogen molecules to maintain high velocity over greater distances, potentially enabling a more rapid dispersion and coverage of larger areas compared to releases at lower altitudes. The decreased temperature at higher altitudes further complicates hydrogen behavior. The interplay of temperature, pressure, and density affects diffusion rates, ignitable mixture formation, and the likelihood of ignition or detonation. These altitude-dependent factors significantly impact hydrogen release dynamics and safety, emphasizing the need for further research. Comprehensive studies are required to fully understand these effects and address the following two key aspects: first, ensuring the safe design and operation of SWITHs across their entire operational altitude range; and, second, replicating these conditions for ground-based experimental testing.
To facilitate the use of hydrogen in SWITHs, developing experimental structural validation test methods is essential to ensure safe working conditions for personnel, instrumentation, and the SWITH itself. However, based on the findings presented in this paper, immediately fulfilling this requirement for large-scale SWITHs, such as those comparable to the Airbus A350-1000 [235], may not be feasible. Instead, a more prudent strategy involves taking smaller, incremental steps towards the larger goal. This gradual approach offers several advantages: it enhances risk management, facilitates knowledge accumulation, and allows researchers to identify and resolve issues at smaller scales before progressing to larger, more complex systems. Additionally, it enables a more cost-effective development process. By following this incremental strategy, researchers can systematically build upon their findings, ensuring a robust and safe development pathway for SWITHs.
The experimental observations presented by Jallais et al. [227], as illustrated in Figure 11, suggest a potential way to mitigate a critical safety concern. To prevent self-ignition or detonation, the working environment could be maintained in a protective gas atmosphere; while acetylene demonstrated enhanced safety properties, helium could be a more suitable option due to its high speed of sound ( 1005 m / s , according to Table 9) and its noble gas characteristics. The inert nature of noble gases, attributed to their fully occupied outer electron shells, makes them highly unreactive. Other noble gases, including neon, argon, krypton, xenon, radon, and oganesson, should also be considered [238,239,240]. The proposed experimental approach involves a structured, incremental process. Initial small-scale experiments should be conducted, systematically increasing either the internal pressure or cylinder volume based on the results. If the proposed inert gases prove ineffective as a surrounding medium, alternative options should be explored. Upon identifying combinations of maximum internal pressure, maximum volume, and surrounding gas that prevent ignition, the research should progress to testing with the addition of external loads. This subsequent phase should be initiated with reduced internal pressure, volume, or a simultaneous decrease in both parameters.
Throughout this critical stage, it is of paramount importance to rigorously investigate various combinations of internal pressure, volume, load type (static or dynamic [241,242,243,244,245]), and external load distribution. Maintaining meticulous records of configurations that can be executed without ignition is vital. This stringent testing strategy ensures a comprehensive understanding of the key variables and helps establish the safest experimental conditions for hydrogen-filled SWITHs. Concurrently with these experimental efforts, simulations should be developed to encapsulate the observed physical behavior. The ultimate objective is to refine experimental testing methods for SWITHs at the scale of large aircraft like the Airbus A350-1000 [235]. This effort must be accompanied by the development of reliable simulation tools to minimize reliance on highly cost-intensive experiments [246]. If substantial discrepancies arise between simulations and experimental results, one possible factor to examine is the influence of the surrounding medium on measurement equipment.
In conclusion, hydrogen self-ignition has been shown to be a tangible concern in real-world applications. This section outlined the primary geometrical factors that can precipitate ignition and explained the modifications necessary to mitigate this risk. Despite the acknowledged risk, the literature supports the feasibility of safe hydrogen applications, provided its properties are thoroughly understood and properly managed. However, caution is crucial when directly applying hydrogen applications from the automotive industry to SWITHs at the scale of large commercial aircraft, such as the Airbus A350-1000. A more judicious strategy involves conducting a series of smaller-scale experimental and numerical studies, incrementally approaching the scale of a large aircraft. This approach enables the systematic accumulation of knowledge and careful risk management when scaling up hydrogen technologies in aviation. A potential pathway to achieving this goal, emphasizing the importance of a gradual, methodical research and development approach, has been outlined.

5. Conclusions

This study investigates the feasibility and safety considerations of using hydrogen as an energy carrier in Wing Structure-Integrated high-pressure Hydrogen Tanks (SWITH), with a particular focus on structural validation tests for aircraft-scale implementations. To establish a foundational understanding of hydrogen’s properties and its viability as an alternative fuel, its fundamental characteristics were examined in detail. This was followed by a theoretical analysis of safety-critical characteristics, highlighting potential hazards such as flammability, detonation risks, material compatibility, and leakage concerns. Additionally, key mitigation strategies were outlined, aiming to support the safe integration of hydrogen into aircraft structures. To validate these insights, findings from numerical simulations and experimental studies available in the literature were reviewed, providing empirical support for the identified safety concerns.
Building upon these insights, a structured framework was proposed for conducting experimental tests on SWITHs, with a strong emphasis on safety. The framework aims to establish testing conditions that minimize risks to the SWITH itself, surrounding personnel, measurement technology, and other critical components required for structural testing. However, in light of these findings, the immediate initiation of large-scale SWITH experiments—without prior risk assessment and gradual scaling—is strongly discouraged. Instead, a structured roadmap should be followed, beginning with small-scale experiments and advancing systematically toward aircraft-sized implementations to ensure safety and reliability. This roadmap underscores the critical need for a thorough experimental and numerical investigation of key variables, including internal pressure, internal volume, load type, and external load distribution. A comprehensive understanding of these factors is essential for the design, operation, and experimental validation of this new aircraft concept. The results of this study align with findings from previous research, which suggest that hydrogen can be utilized safely when its properties are well understood and managed responsibly. A key example of such a responsible approach is the gradual and cautious progression in experimental testing—specifically, refraining from conducting structural tests with high-pressure gaseous hydrogen on large-scale SWITHs at the outset.
The findings of this study have notable implications for the future of aviation, particularly in the context of sustainable fuel alternatives. SWITHs address a key challenge in aviation by acting as a more environmentally friendly and efficient energy solution. However, several technical and safety challenges must be resolved to ensure their successful implementation. Future research should further explore the impact of increasing altitude on the likelihood of ignition transitioning to detonation. Additionally, more advanced investigations into the structural behavior of pressure vessels under external loads are necessary. Developing reliable simulation tools to complement these studies could help reduce reliance on costly experimental testing. Advancements in these areas, alongside an effective certification framework, would pave the way for the safe and scalable adoption of SWITHs in aviation.
Finally, certain key decisions regarding the execution of a preliminary validation test of a small-scale SWITH were guided by the findings of this study. These tests were conducted at IMA Materialforschung und Anwendungstechnik GmbH in Dresden, representing a pioneering endeavor with the potential to shape future structural validation efforts.

Funding

This research was funded by the Federal Ministry for Economic Affairs and Climate Action (grant number: 20M2117B). The Article Processing Charges (APC) were financed by the DFG funded joint publication fund of the TU Dresden, the Medical Faculty Carl Gustav Carus, and the SLUB Dresden. We acknowledge the financial support that made this research possible.

Acknowledgments

Purely for my personal satisfaction, I thank the ONE, who is free of the need for my or anyone else’s acknowledgment, appreciation, or gratitude.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Singh, J.; Srivastawa, K.; Jana, S.; Dixit, C.; S, R. Advancements in Lightweight Materials for Aerospace Structures: A Comprehensive Review. Acceleron Aerosp. J. 2024, 2, 173–183. [Google Scholar] [CrossRef]
  2. Li, S.; Yue, X.; Li, Q.; Peng, H.; Dong, B.; Liu, T.; Yang, H.; Fan, J.; Shu, S.; Qiu, F.; et al. Development and applications of aluminum alloys for aerospace industry. J. Mater. Res. Technol. 2023, 27, 944–983. [Google Scholar] [CrossRef]
  3. Atescan-Yuksek, Y.; Mills, A.; Ayre, D.; Koziol, K.; Salonitis, K. Comparative life cycle assessment of aluminium and CFRP composites: The case of aerospace manufacturing. Int. J. Adv. Manuf. Technol. 2024, 131, 4345–4357. [Google Scholar] [CrossRef]
  4. Chen, Y.; Yuan, M.; Wang, H.; Yu, R.; Hua, L. Progressive optimization on structural design and weight reduction of CFRP key components. Int. J. Lightweight Mater. Manuf. 2023, 6, 59–71. [Google Scholar] [CrossRef]
  5. Acanfora, V.; Sellitto, A.; Russo, A.; Zarrelli, M.; Riccio, A. Experimental investigation on 3D printed lightweight sandwich structures for energy absorption aerospace applications. Aerosp. Sci. Technol. 2023, 137, 108276. [Google Scholar] [CrossRef]
  6. Soni, R.; Verma, R.; Kumar Garg, R.; Sharma, V. A critical review of recent advances in the aerospace materials. Mater. Today Proc. 2024, 113, 180–184. [Google Scholar] [CrossRef]
  7. Okorie, O.; Perveen, A.; Talamona, D.; Kostas, K. Topology Optimization of an Aerospace Bracket: Numerical and Experimental Investigation. Appl. Sci. 2023, 13, 13218. [Google Scholar] [CrossRef]
  8. Ibhadode, O.; Zhang, Z.; Sixt, J.; Nsiempba, K.M.; Orakwe, J.; Martinez-Marchese, A.; Ero, O.; Shahabad, S.I.; Bonakdar, A.; Toyserkani, E. Topology optimization for metal additive manufacturing: Current trends, challenges, and future outlook. Virtual Phys. Prototyp. 2023, 18, 2181192. [Google Scholar] [CrossRef]
  9. Seth, A.; Redonnet, S.; Liem, R.P. MADE: A Multidisciplinary Computational Framework for Aerospace Engineering Education. IEEE Trans. Educ. 2023, 66, 622–631. [Google Scholar] [CrossRef]
  10. Hosseini, S.; Vaziry-Zanjany, M.A.; Ovesy, H.R. A Framework for Aircraft Conceptual Design and Multidisciplinary Optimization. Aerospace 2024, 11, 273. [Google Scholar] [CrossRef]
  11. Leng, J.x.; Wang, Z.g.; Huang, W.; Shen, Y.; An, K. Multidisciplinary Design Optimization Processes for Efficiency Improvement of Aircraft: State-of-the-Art Review. Int. J. Aeronaut. Space Sci. 2024. [Google Scholar] [CrossRef]
  12. Eissele, J.; Lafer, S.; Mejía Burbano, C.; Schließus, J.; Wiedmann, T.; Mangold, J.; Strohmayer, A. Hydrogen-Powered Aviation—Design of a Hybrid-Electric Regional Aircraft for Entry into Service in 2040. Aerospace 2023, 10, 277. [Google Scholar] [CrossRef]
  13. Prewitz, M.; Schwärzer, J.; Bardenhagen, A. Potential analysis of hydrogen storage systems in aircraft design. Int. J. Hydrogen Energy 2023, 48, 25538–25548. [Google Scholar] [CrossRef]
  14. Huete, J.; Nalianda, D.; Pilidis, P. Impact of tank gravimetric efficiency on propulsion system integration for a first-generation hydrogen civil airliner. Aeronaut. J. 2022, 126, 1324–1332. [Google Scholar] [CrossRef]
  15. Cipolla, V.; Zanetti, D.; Abu Salem, K.; Binante, V.; Palaia, G. A Parametric Approach for Conceptual Integration and Performance Studies of Liquid Hydrogen Short–Medium Range Aircraft. Appl. Sci. 2022, 12, 6857. [Google Scholar] [CrossRef]
  16. Adler, E.J.; Martins, J.R. Hydrogen-powered aircraft: Fundamental concepts, key technologies, and environmental impacts. Prog. Aerosp. Sci. 2023, 141, 100922. [Google Scholar] [CrossRef]
  17. Huete, J.; Pilidis, P. Parametric study on tank integration for hydrogen civil aviation propulsion. Int. J. Hydrogen Energy 2021, 46, 37049–37062. [Google Scholar] [CrossRef]
  18. Onorato, G.; Proesmans, P.; Hoogreef, M.F.M. Assessment of hydrogen transport aircraft: Effects of fuel tank integration. CEAS Aeronaut. J. 2022, 13, 813–845. [Google Scholar] [CrossRef]
  19. Baroutaji, A.; Wilberforce, T.; Ramadan, M.; Olabi, A.G. Comprehensive investigation on hydrogen and fuel cell technology in the aviation and aerospace sectors. Renew. Sustain. Energy Rev. 2019, 106, 31–40. [Google Scholar] [CrossRef]
  20. Verstraete, D. Long range transport aircraft using hydrogen fuel. Int. J. Hydrogen Energy 2013, 38, 14824–14831. [Google Scholar] [CrossRef]
  21. Onorato, G. Fuel Tank Integration for Hydrogen Airliners. Master’s Thesis, Delft University of Technology, Delft, The Netherlands, 2021. [Google Scholar]
  22. Verstraete, D. The Potential of Liquid Hydrogen for Long Range Aircraft Propulsion. Ph. D. Thesis, Cranfield University, Wharley End, UK, 2009. [Google Scholar]
  23. Hoelzen, J.; Silberhorn, D.; Zill, T.; Bensmann, B.; Hanke-Rauschenbach, R. Hydrogen-powered aviation and its reliance on green hydrogen infrastructure—Review and research gaps. Int. J. Hydrogen Energy 2022, 47, 3108–3130. [Google Scholar] [CrossRef]
  24. Manigandan, S.; Praveenkumar, T.; Ir Ryu, J.; Nath Verma, T.; Pugazhendhi, A. Role of hydrogen on aviation sector: A review on hydrogen storage, fuel flexibility, flame stability, and emissions reduction on gas turbines engines. Fuel 2023, 352, 129064. [Google Scholar] [CrossRef]
  25. Mueller, T.; Gronau, S. Fostering Macroeconomic Research on Hydrogen-Powered Aviation: A Systematic Literature Review on General Equilibrium Models. Energies 2023, 16, 1439. [Google Scholar] [CrossRef]
  26. Degirmenci, H.; Uludag, A.; Ekici, S.; Karakoc, T.H. Challenges, prospects and potential future orientation of hydrogen aviation and the airport hydrogen supply network: A state-of-art review. Prog. Aerosp. Sci. 2023, 141, 100923. [Google Scholar] [CrossRef]
  27. Gronau, S.; Hoelzen, J.; Mueller, T.; Hanke-Rauschenbach, R. Hydrogen-powered aviation in Germany: A macroeconomic perspective and methodological approach of fuel supply chain integration into an economy-wide dataset. Int. J. Hydrogen Energy 2023, 48, 5347–5376. [Google Scholar] [CrossRef]
  28. Rasul, M.; Hazrat, M.; Sattar, M.; Jahirul, M.; Shearer, M. The future of hydrogen: Challenges on production, storage and applications. Energy Convers. Manag. 2022, 272, 116326. [Google Scholar] [CrossRef]
  29. Capurso, T.; Stefanizzi, M.; Torresi, M.; Camporeale, S. Perspective of the role of hydrogen in the 21st century energy transition. Energy Convers. Manag. 2022, 251, 114898. [Google Scholar] [CrossRef]
  30. Qureshi, F.; Yusuf, M.; Kamyab, H.; Zaidi, S.; Junaid Khalil, M.; Arham Khan, M.; Azad Alam, M.; Masood, F.; Bazli, L.; Chelliapan, S.; et al. Current trends in hydrogen production, storage and applications in India: A review. Sustain. Energy Technol. Assess. 2022, 53, 102677. [Google Scholar] [CrossRef]
  31. Rath, R.; Kumar, P.; Mohanty, S.; Nayak, S.K. Recent advances, unsolved deficiencies, and future perspectives of hydrogen fuel cells in transportation and portable sectors. Int. J. Energy Res. 2019, 43, 8931–8955. [Google Scholar] [CrossRef]
  32. Ustolin, F.; Campari, A.; Taccani, R. An Extensive Review of Liquid Hydrogen in Transportation with Focus on the Maritime Sector. J. Mar. Sci. Eng. 2022, 10, 1222. [Google Scholar] [CrossRef]
  33. Fakhreddine, O.; Gharbia, Y.; Derakhshandeh, J.F.; Amer, A.M. Challenges and Solutions of Hydrogen Fuel Cells in Transportation Systems: A Review and Prospects. World Electr. Veh. J. 2023, 14, 156. [Google Scholar] [CrossRef]
  34. Weigl, J.; Inayati, I.; Zind, E.; Said, H. Pios fuel cell Motorcycle; design, development and test of hydrogen fuel cell powered vehicle. In Proceedings of the 2008 IEEE International Conference on Sustainable Energy Technologies, Singapore, 24–27 November 2008; Volume 127, pp. 1120–1122. [Google Scholar] [CrossRef]
  35. Hwang, J.J.; Chang, W.R. Life-cycle analysis of greenhouse gas emission and energy efficiency of hydrogen fuel cell scooters. Int. J. Hydrogen Energy 2010, 35, 11947–11956. [Google Scholar] [CrossRef]
  36. Hwang, J.J. Review on development and demonstration of hydrogen fuel cell scooters. Renew. Sustain. Energy Rev. 2012, 16, 3803–3815. [Google Scholar] [CrossRef]
  37. Weigl, J.D.; Inayati, I.; Saidi, H. Development of Hydrogen Fuel Cell Motorcycle in South East Asia. ECS Trans. 2011, 30, 289–293. [Google Scholar] [CrossRef]
  38. Dieter Weigl, J.; Henz, M.; Inayati.; Saidi, H. Converted battery-powered electric motorcycle and hydrogen fuel cell-powered electric motorcycle in South East Asia: Development and performance test. In Proceedings of the Joint International Conference on Electric Vehicular Technology and Industrial, Mechanical, Electrical and Chemical Engineering (ICEVT & IMECE), Surakarta, Indonesia, 4–5 November 2015. [Google Scholar] [CrossRef]
  39. Hyundai. Hyundai NEXO: A Hydrogen Powered Car. Available online: https://www.hyundai.com/de/de/modelle/nexo/ausstattung.html (accessed on 30 September 2024).
  40. Hyundai. Hyundai IX35: A Hydrogen Powered Car. Available online: https://ecomento.de/2015/03/26/preis-wasserstoffauto-hyundai-ix35-fuelcell/ (accessed on 30 September 2024).
  41. Toyota. Toyota MIRAI II: A Hydrogen Powered Car. Available online: https://www.toyota.de/neuwagen/mirai (accessed on 30 September 2024).
  42. Honda. Honda Clarity: A Hydrogen Powered Car. Available online: https://hondanews.com/en-US/honda-automobiles/releases/%20release-4f88c507e72a4e7630685979cb04f2cb-2021-clarity-fuel-cell-specifications-features (accessed on 30 September 2024).
  43. Mercedes. Mercedes Benz GLC: A Hydrogen Powered Car. Available online: https://mercedes-benz-media.co.uk/releases/41 (accessed on 30 September 2024).
  44. Barral, K.; Barthélémy, H. Hydrogen high pressure tanks storages: Overview and new trends due to H2 Energy specifications and constraints. WHEC 2006, 20, 1–10. [Google Scholar]
  45. Li, J.C.; Xu, H.; Zhou, K.; Li, J.Q. A review on the research progress and application of compressed hydrogen in the marine hydrogen fuel cell power system. Heliyon 2024, 10, e25304. [Google Scholar] [CrossRef]
  46. Psoma, A.; Sattler, G. Fuel cell systems for submarines: From the first idea to serial production. J. Power Sources 2002, 106, 381–383. [Google Scholar] [CrossRef]
  47. Alstom. Alstom: A Hydrogen Powered Train. Available online: https://www.alstom.com/alstoms-green-traction-solutions-sustainable-solutions-non-electrified-railwaysplasticker.de/preise/ (accessed on 27 September 2024).
  48. De Wagter, C.; Remes, B.; Smeur, E.; van Tienen, F.; Ruijsink, R.; van Hecke, K.; van der Horst, E. The NederDrone: A hybrid lift, hybrid energy hydrogen UAV. Int. J. Hydrogen Energy 2021, 46, 16003–16018. [Google Scholar] [CrossRef]
  49. Yan, Y.; Wang, B.; Wang, C.; Zhao, D.; Xiao, C. Adaptive maximum power point tracking based on Kalman filter for hydrogen fuel cell in hybrid unmanned aerial vehicle applications. Int. J. Hydrogen Energy 2023, 48, 25939–25957. [Google Scholar] [CrossRef]
  50. Kucuk, M.; Sener, R.; Surmen, A. Effectiveness of hydrogen enrichment strategy for Wankel engines in unmanned aerial vehicle applications at various altitudes. Int. J. Hydrogen Energy 2024, 52, 1534–1549. [Google Scholar] [CrossRef]
  51. Tiseira, A.; Novella, R.; Garcia-Cuevas, L.; Lopez-Juarez, M. Concept design and energy balance optimization of a hydrogen fuel cell helicopter for unmanned aerial vehicle and aerotaxi applications. Energy Convers. Manag. 2023, 288, 117101. [Google Scholar] [CrossRef]
  52. Gavrilovic, N.; Mertika, S.; Moschetta, J.M.; Schimpf, J.; Park, G.; Kim, S.Y. Experimental Study on a Liquid Hydrogen Tank for Unmanned Aerial Vehicle Applications. J. Aircr. 2024, 288, 1805–1816. [Google Scholar] [CrossRef]
  53. APUS. Homepage of APUS. Available online: https://group.apus-aero.com/ (accessed on 27 February 2024).
  54. APUS. APUS, Aircraft Manufacturer. 2024. Available online: https://group.apus-aero.com/wp-content/uploads/2023/10/APUS_i-2_Broschuere_20231027_Web.pdf (accessed on 27 September 2024).
  55. Butt, J.A.; Markmiller, J.F.C. Toward Certification of Wing-Structure- Integrated Hydrogen Tanks: A Cross-Industry Review of Relevant Standards. J. Eng. Res. Rep. 2025, 27, 216–239. [Google Scholar] [CrossRef]
  56. Rylko, A.; Van’T Hoff, S.; Lu, L.; Padfield, G.; Podzus, P.; White, M.; Quaranta, G. Rotorcraft Flight Simulation to Support Aircraft Certification: A Review of the State of the Art with an Eye to Future Applications. In Proceedings of the 49th European Rotorcraft Forum (ERF 2023), Bückeburg, Germany, 5–7 September 2023; pp. 1–12. [Google Scholar]
  57. Mauery, T.; Alonso, J.; Cary, A.; Lee, V.; Malecki, R.; Mavriplis, D.; Medic, G.; Schaefer, J.; Slotnick, J. A Guide for Aircraft Certification by Analysis; Technical Report; NASA Langley Research Center: Hampton, VA, USA, 2021.
  58. Narayanan, V.; Chen, T. Research on technology standards: Accomplishment and challenges. Res. Policy 2012, 41, 1375–1406. [Google Scholar] [CrossRef]
  59. Zoo, H.; de Vries, H.J.; Lee, H. Interplay of innovation and standardization: Exploring the relevance in developing countries. Technol. Forecast. Soc. Chang. 2017, 118, 334–348. [Google Scholar] [CrossRef]
  60. Jiang, H.; Zhao, S.; Zhang, Z.J.; Yi, Y. Exploring the mechanism of technology standardization and innovation using the solidification theory of binary eutectic alloy. Technol. Forecast. Soc. Chang. 2018, 135, 217–228. [Google Scholar] [CrossRef]
  61. Morse, E.; Dantan, J.Y.; Anwer, N.; Söderberg, R.; Moroni, G.; Qureshi, A.; Jiang, X.; Mathieu, L. Tolerancing: Managing uncertainty from conceptual design to final product. CIRP Ann. 2018, 67, 695–717. [Google Scholar] [CrossRef]
  62. Klell, M.; Eichlseder, H.; Trattner, A. Wasserstoff in der Fahrzeugtechnik; Springer: Fachmedien Wiesbaden, Germany, 2018. [Google Scholar] [CrossRef]
  63. Klell, M.; Eichlseder, H.; Trattner, A. Hydrogen in Automotive Engineering; Springer: Fachmedien Wiesbaden, Germany, 2023. [Google Scholar] [CrossRef]
  64. Liu, Y.; Wang, Q.; Zhang, J.; Ding, J.; Cheng, Y.; Wang, T.; Li, J.; Hu, F.; Yang, H.B.; Liu, B. Recent Advances in Carbon-Supported Noble-Metal Electrocatalysts for Hydrogen Evolution Reaction: Syntheses, Structures, and Properties. Adv. Energy Mater. 2022, 12, 2200928. [Google Scholar] [CrossRef]
  65. Tarkowski, R.; Uliasz-Misiak, B. Towards underground hydrogen storage: A review of barriers. Renew. Sustain. Energy Rev. 2022, 162, 112451. [Google Scholar] [CrossRef]
  66. Li, Q.; Lu, Y.; Luo, Q.; Yang, X.; Yang, Y.; Tan, J.; Dong, Z.; Dang, J.; Li, J.; Chen, Y.; et al. Thermodynamics and kinetics of hydriding and dehydriding reactions in Mg-based hydrogen storage materials. J. Magnes. Alloy. 2021, 9, 1922–1941. [Google Scholar] [CrossRef]
  67. Kubas, G.J. Fundamentals of H2Binding and Reactivity on Transition Metals Underlying Hydrogenase Function and H2Production and Storage. Chem. Rev. 2007, 107, 4152–4205. [Google Scholar] [CrossRef] [PubMed]
  68. Hwang, J.; Maharjan, K.; Cho, H. A review of hydrogen utilization in power generation and transportation sectors: Achievements and future challenges. Int. J. Hydrogen Energy 2023, 48, 28629–28648. [Google Scholar] [CrossRef]
  69. Madadi Avargani, V.; Zendehboudi, S.; Cata Saady, N.M.; Dusseault, M.B. A comprehensive review on hydrogen production and utilization in North America: Prospects and challenges. Energy Convers. Manag. 2022, 269, 115927. [Google Scholar] [CrossRef]
  70. Pal, D.B.; Singh, A.; Bhatnagar, A. A review on biomass based hydrogen production technologies. Int. J. Hydrogen Energy 2022, 47, 1461–1480. [Google Scholar] [CrossRef]
  71. Tang, D.; Tan, G.L.; Li, G.W.; Liang, J.G.; Ahmad, S.M.; Bahadur, A.; Humayun, M.; Ullah, H.; Khan, A.; Bououdina, M. State-of-the-art hydrogen generation techniques and storage methods: A critical review. J. Energy Storage 2023, 64, 107196. [Google Scholar] [CrossRef]
  72. Jokar, S.; Farokhnia, A.; Tavakolian, M.; Pejman, M.; Parvasi, P.; Javanmardi, J.; Zare, F.; Gonçalves, M.C.; Basile, A. The recent areas of applicability of palladium based membrane technologies for hydrogen production from methane and natural gas: A review. Int. J. Hydrogen Energy 2023, 48, 6451–6476. [Google Scholar] [CrossRef]
  73. Rosen, P.A. Beitrag zur Optimierung von Wasserstoffdruckbehältern; Springer: Fachmedien Wiesbaden, Germany, 2018. [Google Scholar] [CrossRef]
  74. Ishaq, H.; Dincer, I.; Crawford, C. A review on hydrogen production and utilization: Challenges and opportunities. Int. J. Hydrogen Energy 2022, 47, 26238–26264. [Google Scholar] [CrossRef]
  75. Younas, M.; Shafique, S.; Hafeez, A.; Javed, F.; Rehman, F. An Overview of Hydrogen Production: Current Status, Potential, and Challenges. Fuel 2022, 316, 123317. [Google Scholar] [CrossRef]
  76. Dawood, F.; Anda, M.; Shafiullah, G. Hydrogen production for energy: An overview. Int. J. Hydrogen Energy 2020, 45, 3847–3869. [Google Scholar] [CrossRef]
  77. Li, J.Q.; Chen, Y.; Ma, Y.B.; Kwon, J.T.; Xu, H.; Li, J.C. A study on the Joule-Thomson effect of during filling hydrogen in high pressure tank. Case Stud. Therm. Eng. 2023, 41, 102678. [Google Scholar] [CrossRef]
  78. Zhang, C.; Li, Z.; Li, C.; Wu, X.; Sun, L.; Chen, C. Insight into the effects of hydrogen on inside-valve flow and Joule-Thomson characteristics of hydrogen-blended natural gas: A numerical study. Int. J. Hydrogen Energy 2024, 49, 1056–1074. [Google Scholar] [CrossRef]
  79. Li, Z.; Zhang, C.; Li, C.; Wu, X.; Xie, H.; Jiang, L. Thermodynamic evaluation of the effects of hydrogen blending on the Joule-Thomson characteristics of hydrogen-blended natural gas. J. Clean. Prod. 2023, 406, 137074. [Google Scholar] [CrossRef]
  80. Sigloch, H. Technische Fluidmechanik; Springer: Berlin/Heidelberg, Germany, 2022. [Google Scholar] [CrossRef]
  81. Usman, M.R. Hydrogen storage methods: Review and current status. Renew. Sustain. Energy Rev. 2022, 167, 112743. [Google Scholar] [CrossRef]
  82. Yang, M.; Hunger, R.; Berrettoni, S.; Sprecher, B.; Wang, B. A review of hydrogen storage and transport technologies. Clean Energy 2023, 7, 190–216. [Google Scholar] [CrossRef]
  83. Panigrahi, P.K.; Chandu, B.; Motapothula, M.R.; Puvvada, N. Potential Benefits, Challenges and Perspectives of Various Methods and Materials Used for Hydrogen Storage. Energy Fuels 2024, 38, 2630–2653. [Google Scholar] [CrossRef]
  84. Drawer, C.; Lange, J.; Kaltschmitt, M. Metal hydrides for hydrogen storage—Identification and evaluation of stationary and transportation applications. J. Energy Storage 2024, 77, 109988. [Google Scholar] [CrossRef]
  85. Nivedhitha, K.; Beena, T.; Banapurmath, N.; Umarfarooq, M.; Ramasamy, V.; Soudagar, M.E.M.; Ağbulut, Ü. Advances in hydrogen storage with metal hydrides: Mechanisms, materials, and challenges. Int. J. Hydrogen Energy 2024, 61, 1259–1273. [Google Scholar] [CrossRef]
  86. Boateng, E.; Thiruppathi, A.R.; Hung, C.K.; Chow, D.; Sridhar, D.; Chen, A. Functionalization of graphene-based nanomaterials for energy and hydrogen storage. Electrochim. Acta 2023, 452, 142340. [Google Scholar] [CrossRef]
  87. Yan, Y.; Zhang, J.; Li, G.; Zhou, W.; Ni, Z. Review on linerless type V cryo-compressed hydrogen storage vessels: Resin toughening and hydrogen-barrier properties control. Renew. Sustain. Energy Rev. 2024, 189, 114009. [Google Scholar] [CrossRef]
  88. Aceves, S.M.; Espinosa-Loza, F.; Ledesma-Orozco, E.; Ross, T.O.; Weisberg, A.H.; Brunner, T.C.; Kircher, O. High-density automotive hydrogen storage with cryogenic capable pressure vessels. Int. J. Hydrogen Energy 2010, 35, 1219–1226. [Google Scholar] [CrossRef]
  89. Jaramillo, D.E.; Moreno-Blanco, J.; Aceves, S.M. Evaluation of cryo-compressed hydrogen for heavy-duty trucks. Int. J. Hydrogen Energy 2024, 87, 928–938. [Google Scholar] [CrossRef]
  90. Faye, O.; Szpunar, J.; Eduok, U. A critical review on the current technologies for the generation, storage, and transportation of hydrogen. Int. J. Hydrogen Energy 2022, 47, 13771–13802. [Google Scholar] [CrossRef]
  91. Klopčič, N.; Grimmer, I.; Winkler, F.; Sartory, M.; Trattner, A. A review on metal hydride materials for hydrogen storage. J. Energy Storage 2023, 72, 108456. [Google Scholar] [CrossRef]
  92. Hassan, Q.; Algburi, S.; Sameen, A.Z.; Jaszczur, M.; Salman, H.M. Hydrogen as an energy carrier: Properties, storage methods, challenges, and future implications. Environ. Syst. Decis. 2023, 44, 327–350. [Google Scholar] [CrossRef]
  93. Dryer, F.L.; Chaos, M.; Zhao, Z.; Stein, J.N.; Alpert, J.Y.; Homer, C.J. Spontaneous ignition of pressurized releases of hydrogen and natural gas into air. Combust. Sci. Technol. 2007, 179, 663–694. [Google Scholar] [CrossRef]
  94. Thomas, C.E. Direct-Hydrogen-Fueled Proton-Exchange-Membrane Fuel Cell System for Transportation Applications; Hydrogen Vehicle Safety Report; Ford Motor Co.: Dearborn, MI, USA, 1997. [Google Scholar] [CrossRef]
  95. Farhana, K.; Shadate Faisal Mahamude, A.; Kadirgama, K. Comparing hydrogen fuel cost of production from various sources—A competitive analysis. Energy Convers. Manag. 2024, 302, 118088. [Google Scholar] [CrossRef]
  96. Yusaf, T.; Faisal Mahamude, A.S.; Kadirgama, K.; Ramasamy, D.; Farhana, K.; A. Dhahad, H.; Abu Talib, A.R. Sustainable hydrogen energy in aviation—A narrative review. Int. J. Hydrogen Energy 2024, 52, 1026–1045. [Google Scholar] [CrossRef]
  97. Yao, J.; Wu, Z.; Wang, H.; Yang, F.; Ren, J.; Zhang, Z. Application-oriented hydrolysis reaction system of solid-state hydrogen storage materials for high energy density target: A review. J. Energy Chem. 2022, 74, 218–238. [Google Scholar] [CrossRef]
  98. Prewitz, M.; Bardenhagen, A.; Beck, R. Hydrogen as the fuel of the future in aircrafts—Challenges and opportunities. Int. J. Hydrogen Energy 2020, 45, 25378–25385. [Google Scholar] [CrossRef]
  99. Tashie-Lewis, B.C.; Nnabuife, S.G. Hydrogen Production, Distribution, Storage and Power Conversion in a Hydrogen Economy—A Technology Review. Chem. Eng. J. Adv. 2021, 8, 100172. [Google Scholar] [CrossRef]
  100. Kubilay Karayel, G.; Javani, N.; Dincer, I. A comprehensive assessment of energy storage options for green hydrogen. Energy Convers. Manag. 2023, 291, 117311. [Google Scholar] [CrossRef]
  101. Yanxing, Z.; Maoqiong, G.; Yuan, Z.; Xueqiang, D.; Jun, S. Thermodynamics analysis of hydrogen storage based on compressed gaseous hydrogen, liquid hydrogen and cryo-compressed hydrogen. Int. J. Hydrogen Energy 2019, 44, 16833–16840. [Google Scholar] [CrossRef]
  102. Wang, Z.; Wang, Y.; Afshan, S.; Hjalmarsson, J. A review of metallic tanks for H2 storage with a view to application in future green shipping. Int. J. Hydrogen Energy 2021, 46, 6151–6179. [Google Scholar] [CrossRef]
  103. Jella, S.; Gupta, R.B.; Gupta, A.; Bhadoria, S.K. Computational study of type IV hydrogen tanks. In Proceedings of the 14th International Conference on Materials Processing and Characterization, Hyderabad, India, 24–26 March 2024; Volume 3007, p. 030036. [Google Scholar] [CrossRef]
  104. Cheng, Q.; Zhang, R.; Shi, Z.; Lin, J. Review of common hydrogen storage tanks and current manufacturing methods for aluminium alloy tank liners. Int. J. Lightweight Mater. Manuf. 2024, 7, 269–284. [Google Scholar] [CrossRef]
  105. Rivard, E.; Trudeau, M.; Zaghib, K. Hydrogen Storage for Mobility: A Review. Materials 2019, 12, 1973. [Google Scholar] [CrossRef]
  106. Air, A.; Shamsuddoha, M.; Gangadhara Prusty, B. A review of Type V composite pressure vessels and automated fibre placement based manufacturing. Compos. Part Eng. 2023, 253, 110573. [Google Scholar] [CrossRef]
  107. Air, A.; Oromiehie, E.; Prusty, B.G. Design and manufacture of a Type V composite pressure vessel using automated fibre placement. Compos. Part Eng. 2023, 266, 111027. [Google Scholar] [CrossRef]
  108. Shin, H.K.; Ha, S.K. A Review on the Cost Analysis of Hydrogen Gas Storage Tanks for Fuel Cell Vehicles. Energies 2023, 16, 5233. [Google Scholar] [CrossRef]
  109. Shin, H.K.; Ha, S.K. Techno-economic analysis of type III and IV composite hydrogen storage tanks for fuel cell vehicles. Adv. Compos. Mater. 2024, 33, 527–559. [Google Scholar] [CrossRef]
  110. Mori, D.; Hirose, K. Recent challenges of hydrogen storage technologies for fuel cell vehicles. Int. J. Hydrogen Energy 2009, 34, 4569–4574. [Google Scholar] [CrossRef]
  111. Mahmoud, M. Development of a New Correlation of Gas Compressibility Factor (Z-Factor) for High Pressure Gas Reservoirs. J. Energy Resour. Technol. 2013, 136, 012903. [Google Scholar] [CrossRef]
  112. Lemmon, E.W.; Huber, M.L.; Leachman, J.W. Revised Standardized Equation for Hydrogen Gas Densities for Fuel Consumption Applications. J. Res. Natl. Inst. Stand. Technol. 2008, 113, 341. [Google Scholar] [CrossRef] [PubMed]
  113. Stephan, P.; Schaber, K.; Stephan, K.; Mayinger, F. Thermodynamik: Grundlagen und Technische Anwendungen Band 1: Einstoffsysteme; Springer: Berlin/Heidelberg, Germany, 2013. [Google Scholar] [CrossRef]
  114. Zheng, J.; Zhang, X.; Xu, P.; Gu, C.; Wu, B.; Hou, Y. Standardized equation for hydrogen gas compressibility factor for fuel consumption applications. Int. J. Hydrogen Energy 2016, 41, 6610–6617. [Google Scholar] [CrossRef]
  115. Chen, H.; Zheng, J.; Xu, P.; Li, L.; Liu, Y.; Bie, H. Study on real-gas equations of high pressure hydrogen. Int. J. Hydrogen Energy 2010, 35, 3100–3104. [Google Scholar] [CrossRef]
  116. Davarnejad, R.; Rahimi, B.; Baghban, S.N.; Khansary, M.A.; Sani, A.H. Development of a thermodynamic model for hydrogen and hydrogen containing mixtures. Fluid Phase Equilibria 2014, 382, 1–9. [Google Scholar] [CrossRef]
  117. Werkle, H. Finite Elements in Structural Analysis: Theoretical Concepts and Modeling Procedures in Statics and Dynamics of Structures; Springer International Publishing: Cham, Switzerland, 2021. [Google Scholar] [CrossRef]
  118. Langtangen, H.P.; Logg, A. Solving PDEs in Python; Springer International Publishing: Cham, Switzerland, 2016. [Google Scholar] [CrossRef]
  119. Langtangen, H.P.; Mardal, K.A. Introduction to Numerical Methods for Variational Problems; Springer International Publishing: Cham, Switzerland, 2019. [Google Scholar] [CrossRef]
  120. Kim, K.; Erickson, A.; Lambert, A.; Bruder, G.; Welch, G. Effects of Dark Mode on Visual Fatigue and Acuity in Optical See-Through Head-Mounted Displays. In Proceedings of the Symposium on Spatial User Interaction, SUI’19, New Orleans, LA, USA, 19–20 October 2019; pp. 1–9. [Google Scholar] [CrossRef]
  121. Chen, C.H.; Zhai, W. The Effects of Information Layout, Display Mode, and Gender Difference on the User Interface Design of Mobile Shopping Applications. IEEE Access 2014, 11, 47024–47039. [Google Scholar] [CrossRef]
  122. Tunes, M.A.; Uggowitzer, P.J.; Dumitraschkewitz, P.; Willenshofer, P.; Samberger, S.; da Silva, F.C.; Schön, C.G.; Kremmer, T.M.; Antrekowitsch, H.; Djukic, M.B.; et al. Limitations of Hydrogen Detection After 150 Years of Research on Hydrogen Embrittlement. Adv. Eng. Mater. 2024, 26, 2400776. [Google Scholar] [CrossRef]
  123. Ma, Y.; Liang, T.; Qiao, S.; Liu, X.; Lang, Z. Highly Sensitive and Fast Hydrogen Detection Based on Light-Induced Thermoelastic Spectroscopy. Ultrafast Sci. 2023, 3, 24. [Google Scholar] [CrossRef]
  124. Darmadi, I.; Östergren, I.; Lerch, S.; Lund, A.; Moth-Poulsen, K.; Müller, C.; Langhammer, C. Bulk-Processed Plasmonic Plastic Nanocomposite Materials for Optical Hydrogen Detection. Accounts Chem. Res. 2023, 56, 1850–1861. [Google Scholar] [CrossRef]
  125. European Union Aviation Safety Agency. Certification Specifications, Acceptable Means of Compliance and Guidance Material for Sailplanes and Powered Sailplanes (CS-22); Certification Specification; European Union Aviation Safety Agency: Cologne, Germany, 2008.
  126. European Union Aviation Safety Agency. Certification Specifications for Normal, Utility, Aerobatic, and Commuter Category Aeroplanes CS-23; Certification Specification; European Union Aviation Safety Agency: Cologne, Germany, 2009.
  127. European Union Aviation Safety Agency. Certification Specifications and Acceptable Means of Compliance for Large Aeroplanes (CS-25); Certification Specification; European Union Aviation Safety Agency: Cologne, Germany, 2010.
  128. European Union Aviation Safety Agency. Certification Specifications, Acceptable Means of Compliance and Guidance Material for Small Rotorcraft CS-27; Certification Specification; European Union Aviation Safety Agency: Cologne, Germany, 2023.
  129. European Union Aviation Safety Agency. Certification Specifications, Acceptable Means of Compliance and Guidance Material for Large Rotorcraft CS-29; Certification Specification; European Union Aviation Safety Agency: Cologne, Germany, 2007.
  130. SAE International. Liquid Hydrogen Storage for Aviation AS6679, WIP 2019-11-20. Available online: https://www.sae.org/standards/content/as6679/ (accessed on 6 September 2024).
  131. SAE International. Gaseous Hydrogen Storage for General Aviation AS7373, WIP 2021-07-06. Available online: https://www.sae.org/standards/content/as7373/ (accessed on 6 September 2024).
  132. European Union. Regulation (EC) No 79/2009 of the European Parliament and of the Council of 14 January 2009 on Type-Approval of Hydrogen-Powered Motor Vehicles, and Amending Directive 2007/46/EC (Text with EEA Relevance), Document 32009R0079; Regulation; European Union: Brussels, Belgium, 2009.
  133. European Union. Commission Regulation (EU) No 406/2010 of 26 April 2010 Implementing Regulation (EC) No 79/2009 of the European Parliament and of the Council on Type-Approval of Hydrogen-Powered Motor Vehicles (Text with EEA Relevance), Document 32010R0406; Regulation; European Union: Brussels, Belgium, 2010.
  134. European Union. Regulation No 134 of the Economic Commission for Europe of the United Nations (UN/ECE)—Uniform Provisions Concerning the Approval of Motor Vehicles and Their Components with Regard to the Safety-Related Performance of Hydrogen-Fuelled Vehicles (HFCV) [2019/795]; Regulation; European Union: Brussels, Belgium, 2019.
  135. Global Technical Regulation. UN GTR No. 13—Global Technical Regulation Concerning the Hydrogen and Fuel Cell Vehicles; Regulation; Global Technical Regulation: Brussels, Belgium, 2013. [Google Scholar]
  136. American National Standards Institute. CSA/ANSI HGV 2-2014 (R2019), ANSI HGV 2-2014—Compressed Hydrogen Gas Vehicle Fuel Containers; Regulation; American National Standards Institute: Washington, DA, USA, 2014. [Google Scholar]
  137. SAE International. Standard for Fuel Systems in Fuel Cell and Other Hydrogen Vehicles J2579_202301; Regulation; SAE International: Warrendale, PA, USA, 2013. [Google Scholar]
  138. Iijima, T.; Abe, T.; Itoga, H. Development of material testing equipment in high pressure gaseous hydrogen and international collaborative work of a testing method for a hydrogen society—Toward contribution to international standardization. Synth. Engl. Ed. 2015, 8, 61–69. [Google Scholar] [CrossRef]
  139. Japan Automobile Standards Internationalization Center (JASIC). Japan—JARI S001(2004) Technical Standard for Containers of Compressed Hydrogen Vehicle Fuel Devices. Available online: https://unece.org/DAM/trans/doc/2008/wp29grsp/SGS-3-09e.pdf (accessed on 6 September 2024).
  140. ISO/TS 15869:2009; Gaseous Hydrogen and Hydrogen Blends—Land Vehicle Fuel Tanks. Regulation; International Organization for Standardization: Geneva, Switzerland, 2009.
  141. ISO 19881:2018; Gaseous Hydrogen — Land Vehicle Fuel Containers. Regulation; International Organization for Standardization: Geneva, Switzerland, 2018.
  142. ISO/DIS 19881; Gaseous Hydrogen—Land Vehicle Fuel Containers. Regulation; International Organization for Standardization: Geneva, Switzerland, 1988.
  143. ISO 11119-3; Gas Cylinders—Design, Construction and Testing of Refillable Composite Gas Cylinders and Tubes—Part 3: Fully Wrapped Fibre REINFORCED Composite Gas Cylinders and Tubes up to 450 L with Non-Load-Sharing Metallic or Non-Metallic Liners or Without Liners. International Organization for Standardization: Geneva, Switzerland, 2020.
  144. Cheikhravat, H.; Chaumeix, N.; Bentaib, A.; Paillard, C.E. Flammability Limits of Hydrogen-Air Mixtures. Nucl. Technol. 2012, 178, 5–16. [Google Scholar] [CrossRef]
  145. Crowl, D.A.; Jo, Y.D. The hazards and risks of hydrogen. J. Loss Prev. Process Ind. 2007, 20, 158–164. [Google Scholar] [CrossRef]
  146. Sánchez, A.L.; Williams, F.A. Recent advances in understanding of flammability characteristics of hydrogen. Prog. Energy Combust. Sci. 2014, 41, 1–55. [Google Scholar] [CrossRef]
  147. Zhang, K.; Luo, T.; Li, Y.; Zhang, T.; Li, X.; Zhang, Z.; Shang, S.; Zhou, Y.; Zhang, C.; Chen, X.; et al. Effect of ignition, initial pressure and temperature on the lower flammability limit of hydrogen/air mixture. Int. J. Hydrogen Energy 2022, 47, 15107–15119. [Google Scholar] [CrossRef]
  148. Bane, S.; Ziegler, J.; Boettcher, P.; Coronel, S.; Shepherd, J. Experimental investigation of spark ignition energy in kerosene, hexane, and hydrogen. J. Loss Prev. Process Ind. 2013, 26, 290–294. [Google Scholar] [CrossRef]
  149. Dimitriou, P.; Tsujimura, T. A review of hydrogen as a compression ignition engine fuel. Int. J. Hydrogen Energy 2017, 42, 24470–24486. [Google Scholar] [CrossRef]
  150. Fernández-Tarrazo, E.; Gómez-Miguel, R.; Sánchez-Sanz, M. Minimum ignition energy of hydrogen–ammonia blends in air. Fuel 2023, 337, 127128. [Google Scholar] [CrossRef]
  151. Yang, F.; Wang, T.; Deng, X.; Dang, J.; Huang, Z.; Hu, S.; Li, Y.; Ouyang, M. Review on hydrogen safety issues: Incident statistics, hydrogen diffusion, and detonation process. Int. J. Hydrogen Energy 2021, 46, 31467–31488. [Google Scholar] [CrossRef]
  152. Gao, Y.; Zhang, B.; Ng, H.D.; Lee, J.H. An experimental investigation of detonation limits in hydrogen–oxygen–argon mixtures. Int. J. Hydrogen Energy 2016, 41, 6076–6083. [Google Scholar] [CrossRef]
  153. Zhang, B. The influence of wall roughness on detonation limits in hydrogen–oxygen mixture. Combust. Flame 2016, 169, 333–339. [Google Scholar] [CrossRef]
  154. Sullivan, J.H.; Feber, R.C.; Starner, J.W. Mechanism and types of explosive behavior in hydrogen–fluorine systems. J. Chem. Phys. 1975, 62, 1714–1725. [Google Scholar] [CrossRef]
  155. Gustin, J.L. Safety of chlorine production and chlorination processes. Chem. Health Saf. 2005, 12, 5–16. [Google Scholar] [CrossRef]
  156. Ono, R.; Nifuku, M.; Fujiwara, S.; Horiguchi, S.; Oda, T. Minimum ignition energy of hydrogen–air mixture: Effects of humidity and spark duration. J. Electrost. 2007, 65, 87–93. [Google Scholar] [CrossRef]
  157. Cirrone, D.; Makarov, D.; Proust, C.; Molkov, V. Minimum ignition energy of hydrogen-air mixtures at ambient and cryogenic temperatures. Int. J. Hydrogen Energy 2023, 48, 16530–16544. [Google Scholar] [CrossRef]
  158. Wen, J.X.; Marono, M.; Moretto, P.; Reinecke, E.A.; Sathiah, P.; Studer, E.; Vyazmina, E.; Melideo, D. Statistics, lessons learned and recommendations from analysis of HIAD 2.0 database. Int. J. Hydrogen Energy 2022, 47, 17082–17096. [Google Scholar] [CrossRef]
  159. Mirza, N.R.; Degenkolbe, S.; Witt, W. Analysis of hydrogen incidents to support risk assessment. Int. J. Hydrogen Energy 2011, 36, 12068–12077. [Google Scholar] [CrossRef]
  160. Weiner, S.; Fassbender, L.; Quick, K. Using hydrogen safety best practices and learning from safety events. Int. J. Hydrogen Energy 2011, 36, 2729–2735. [Google Scholar] [CrossRef]
  161. Campari, A.; Nakhal Akel, A.J.; Ustolin, F.; Alvaro, A.; Ledda, A.; Agnello, P.; Moretto, P.; Patriarca, R.; Paltrinieri, N. Lessons learned from HIAD 2.0: Inspection and maintenance to avoid hydrogen-induced material failures. Comput. Chem. Eng. 2023, 173, 108199. [Google Scholar] [CrossRef]
  162. Stolzenburg, K.; Tsatsami, V.; Grubel, H. Lessons learned from infrastructure operation in the CUTE project. Int. J. Hydrogen Energy 2009, 34, 7114–7124. [Google Scholar] [CrossRef]
  163. Alfasfos, R.; Sillman, J.; Soukka, R. Lessons learned and recommendations from analysis of hydrogen incidents and accidents to support risk assessment for the hydrogen economy. Int. J. Hydrogen Energy 2024, 60, 1203–1214. [Google Scholar] [CrossRef]
  164. Schefer, R.W.; Kulatilaka, W.D.; Patterson, B.D.; Settersten, T.B. Visible emission of hydrogen flames. Combust. Flame 2009, 156, 1234–1241. [Google Scholar] [CrossRef]
  165. Salaudeen, S.; Arku, P.; Dutta, A. Gasification of Plastic Solid Waste and Competitive Technologies. In Plastics to Energy; Elsevier: Amsterdam, The Netherlands, 2019; pp. 269–293. [Google Scholar] [CrossRef]
  166. Liang, W.; Liu, J.; Law, C.K. On explosion limits of H2/CO/O2 mixtures. Combust. Flame 2017, 179, 130–137. [Google Scholar] [CrossRef]
  167. Li, S.; Liang, W.; Yao, Q.; Law, C.K. An analysis of the ignition limits of premixed hydrogen/oxygen by heated nitrogen in counterflow. Combust. Flame 2018, 198, 230–239. [Google Scholar] [CrossRef]
  168. Liang, W.; Wang, Y.; Law, C.K. Role of ozone doping in the explosion limits of hydrogen-oxygen mixtures: Multiplicity and catalyticity. Combust. Flame 2019, 205, 7–10. [Google Scholar] [CrossRef]
  169. Liu, J.; Yu, R.; Liang, W.; Law, C.K. On explosion-limit regime diagram of H2 and C1 to C3 alkanes with unified pivot state. Combust. Flame 2023, 251, 112705. [Google Scholar] [CrossRef]
  170. Liu, J.; Wang, J.; Zhang, N.; Zhao, H. On the explosion limit of syngas with CO2 and H2O additions. Int. J. Hydrogen Energy 2018, 43, 3317–3329. [Google Scholar] [CrossRef]
  171. Zhou, C.; Liang, W.; Chen, Z. On Explosion Limits of Ammonia–Oxygen Mixtures with Hydrogen Addition: Sensitivity and Nonmonotonicity. Energy Fuels 2021, 35, 14035–14041. [Google Scholar] [CrossRef]
  172. Li, J.; Liang, W.; Han, W. A comparative study of chlorine and bromine species addition on the explosion limits of hydrogen-oxygen mixtures. Int. J. Hydrogen Energy 2023, 48, 32125–32136. [Google Scholar] [CrossRef]
  173. Li, X.; Ma, X.; Zhang, J.; Akiyama, E.; Wang, Y.; Song, X. Review of Hydrogen Embrittlement in Metals: Hydrogen Diffusion, Hydrogen Characterization, Hydrogen Embrittlement Mechanism and Prevention. Acta Metall. Sin. 2020, 33, 759–773. [Google Scholar] [CrossRef]
  174. Jia, G.; Lei, M.; Li, M.; Xu, W.; Li, R.; Lu, Y.; Cai, M. Hydrogen embrittlement in hydrogen-blended natural gas transportation systems: A review. Int. J. Hydrogen Energy 2023, 48, 32137–32157. [Google Scholar] [CrossRef]
  175. Campari, A.; Ustolin, F.; Alvaro, A.; Paltrinieri, N. A review on hydrogen embrittlement and risk-based inspection of hydrogen technologies. Int. J. Hydrogen Energy 2023, 48, 35316–35346. [Google Scholar] [CrossRef]
  176. Meda, U.S.; Bhat, N.; Pandey, A.; Subramanya, K.; Lourdu Antony Raj, M. Challenges associated with hydrogen storage systems due to the hydrogen embrittlement of high strength steels. Int. J. Hydrogen Energy 2023, 48, 17894–17913. [Google Scholar] [CrossRef]
  177. Gong, P.; Turk, A.; Nutter, J.; Yu, F.; Wynne, B.; Rivera-Diaz-del Castillo, P.; Mark Rainforth, W. Hydrogen embrittlement mechanisms in advanced high strength steel. Acta Mater. 2023, 2022, 117488. [Google Scholar] [CrossRef]
  178. Du, Z.; Liu, C.; Zhai, J.; Guo, X.; Xiong, Y.; Su, W.; He, G. A Review of Hydrogen Purification Technologies for Fuel Cell Vehicles. Catalysts 2021, 11, 393. [Google Scholar] [CrossRef]
  179. Nashchekin, M.D.; Minko, M.V.; Morgunova, S.B.; Minko, K.B. Enhancement of heat- and mass-transfer processes in a metal-hydride flow-through hydrogen-purification reactor. Int. J. Hydrogen Energy 2020, 45, 25013–25029. [Google Scholar] [CrossRef]
  180. Su, Y.; Lv, H.; Zhou, W.; Zhang, C. Review of the Hydrogen Permeability of the Liner Material of Type IV On-Board Hydrogen Storage Tank. World Electr. Veh. J. 2021, 12, 130. [Google Scholar] [CrossRef]
  181. McLean, D. Understanding Aerodynamics, Reprinted with Corrections ed.; Aerospace Series; Wiley: Hoboken, NJ, USA, 2014. [Google Scholar]
  182. Vos, R.; Farokhi, S. Introduction to Transonic Aerodynamics; Springer: Dordrecht, The Netherlands, 2015. [Google Scholar] [CrossRef]
  183. Liu, P. Aerodynamics; Springer Nature: Singapore, 2022. [Google Scholar] [CrossRef]
  184. Jiang, Y.; Pan, X.; Cai, Q.; Klymenko, O.V.; Hua, M.; Zhang, T.; Wang, Z.; Wang, Q.; Yu, A.; Jiang, J. Effects of the partially open inlet on shock waves and spontaneous ignition during the leakage of hydrogen. Process Saf. Environ. Prot. 2022, 168, 1089–1100. [Google Scholar] [CrossRef]
  185. Xu, B.; Hima, L.E.; Wen, J.; Tam, V. Numerical study of spontaneous ignition of pressurized hydrogen release into air. Int. J. Hydrogen Energy 2009, 34, 5954–5960. [Google Scholar] [CrossRef]
  186. Goroshin, S.; Frost, D.; Levine, J.; Yoshinaka, A.; Zhang, F. Optical Pyrometry of Fireballs of Metalized Explosives. Propellants Explos. Pyrotech. Int. J. Deal. Sci. Technol. Asp. Energetic Mater. 2006, 31, 169–181. [Google Scholar] [CrossRef]
  187. Frost, D.L.; Clemenson, J.; Goroshin, S.; Zhang, F.; Soo, M. Thermocouple Temperature Measurements in Metalized Explosive Fireballs. Propellants Explos. Pyrotech. Int. J. Deal. Sci. Technol. Asp. Energetic Mater. 2021, 46, 899–911. [Google Scholar] [CrossRef]
  188. Lebel, L.S.; Brousseau, P.; Erhardt, L.; Andrews, W.S. Measurements of the Temperature Inside an Explosive Fireball. J. Appl. Mech. 2013, 80, 031702. [Google Scholar] [CrossRef]
  189. Cashdollar, K.L.; Zlochower, I.A. Explosion temperatures and pressures of metals and other elemental dust clouds. J. Loss Prev. Process Ind. 2007, 20, 337–348. [Google Scholar] [CrossRef]
  190. Kim, Y.R.; Lee, H.J.; Kim, S.; Jeung, I.S. A flow visualization study on self-ignition of high pressure hydrogen gas released into a tube. Proc. Combust. Inst. 2013, 34, 2057–2064. [Google Scholar] [CrossRef]
  191. Sasoh, A. Compressible Fluid Dynamics and Shock Waves; Springer: Singapore, 2020. [Google Scholar] [CrossRef]
  192. Stauffer, E.; Dolan, J.A.; Newman, R. Chemistry and Physics of Fire and Liquid Fuels. In Fire Debris Analysis; Elsevier: Amsterdam, The Netherlands, 2008; pp. 85–129. [Google Scholar] [CrossRef]
  193. Astbury, G.; Hawksworth, S.J. Spontaneous ignition of hydrogen leaks: A review of postulated mechanisms. Int. J. Hydrogen Energy 2007, 32, 2178–2185. [Google Scholar] [CrossRef]
  194. Zhou, S.; Luo, Z.; Wang, T.; He, M.; Li, R.; Su, B. Research progress on the self-ignition of high-pressure hydrogen discharge: A review. Int. J. Hydrogen Energy 2022, 47, 9460–9476. [Google Scholar] [CrossRef]
  195. Qiu, H.; Zhou, R.; Li, X.; Xie, Y.; Fan, M.; Li, J.; Huang, H. A review on spontaneous ignition mechanism of pressurized hydrogen released through tubes. Int. J. Hydrogen Energy 2024, 86, 613–637. [Google Scholar] [CrossRef]
  196. Wolanski, P. Investigation into the mechanism of the diffusion ignition of a combustible gas flowing into an oxidizing atmosphere. In Proceedings of the Fourteenth Symposium (International) on Combustion, University Park, PE, USA, 20–25 August 1973. [Google Scholar]
  197. Mogi, T.; Kim, D.; Shiina, H.; Horiguchi, S. Self-ignition and explosion during discharge of high-pressure hydrogen. J. Loss Prev. Process Ind. 2008, 21, 199–204. [Google Scholar] [CrossRef]
  198. Golub, V.; Baklanov, D.; Golovastov, S.; Ivanov, M.; Laskin, I.; Saveliev, A.; Semin, N.; Volodin, V. Mechanisms of high-pressure hydrogen gas self-ignition in tubes. J. Loss Prev. Process Ind. 2008, 21, 185–198. [Google Scholar] [CrossRef]
  199. Mogi, T.; Wada, Y.; Ogata, Y.; Koichi Hayashi, A. Self-ignition and flame propagation of high-pressure hydrogen jet during sudden discharge from a pipe. Int. J. Hydrogen Energy 2009, 34, 5810–5816. [Google Scholar] [CrossRef]
  200. Lee, H.J.; Kim, Y.R.; Kim, S.H.; Jeung, I.S. Experimental investigation on the self-ignition of pressurized hydrogen released by the failure of a rupture disk through tubes. Proc. Combust. Inst. 2011, 33, 2351–2358. [Google Scholar] [CrossRef]
  201. Kim, S.; Lee, H.J.; Park, J.H.; Jeung, I.S. Effects of a wall on the self-ignition patterns and flame propagation of high-pressure hydrogen release through a tube. Proc. Combust. Inst. 2013, 34, 2049–2056. [Google Scholar] [CrossRef]
  202. Kitabayashi, N.; Wada, Y.; Mogi, T.; Saburi, T.; Hayashi, A. Experimental study on high pressure hydrogen jets coming out of tubes of 0.1–4.2 m in length. Int. J. Hydrogen Energy 2013, 38, 8100–8107. [Google Scholar] [CrossRef]
  203. Grune, J.; Sempert, K.; Kuznetsov, M.; Jordan, T. Experimental investigation of flame and pressure dynamics after spontaneous ignition in tube geometry. Int. J. Hydrogen Energy 2014, 39, 20396–20403. [Google Scholar] [CrossRef]
  204. Gong, L.; Duan, Q.; Jiang, L.; Jin, K.; Sun, J. Experimental study of pressure dynamics, spontaneous ignition and flame propagation during hydrogen release from high-pressure storage tank through 15 mm diameter tube and exhaust chamber connected to atmosphere. Fuel 2016, 182, 419–427. [Google Scholar] [CrossRef]
  205. Duan, Q.; Xiao, H.; Gao, W.; Gong, L.; Wang, Q.; Sun, J. Experimental study on spontaneous ignition and flame propagation of high-pressure hydrogen release via a tube into air. Fuel 2016, 181, 811–819. [Google Scholar] [CrossRef]
  206. Jiang, Y.; Pan, X.; Yan, W.; Wang, Z.; Wang, Q.; Hua, M.; Jiang, J. Pressure dynamics, self-ignition, and flame propagation of hydrogen jet discharged under high pressure. Int. J. Hydrogen Energy 2019, 44, 22661–22670. [Google Scholar] [CrossRef]
  207. Wang, Z.; Pan, X.; Wang, Q.; Jiang, Y.; Xu, X.; Yan, W.; Jiang, J. Experimental study on spontaneous ignition and flame propagation of high-pressure hydrogen release through tubes. Int. J. Hydrogen Energy 2019, 44, 22584–22597. [Google Scholar] [CrossRef]
  208. Jin, K.; Yang, S.; Gong, L.; Han, Y.; Yang, X.; Gao, Y.; Zhang, Y. Numerical study on the spontaneous ignition of pressurized hydrogen during its sudden release into the tube with varying lengths and diameters. J. Loss Prev. Process Ind. 2021, 72, 104592. [Google Scholar] [CrossRef]
  209. Butt, J.A. Development of a Module for Mission Analysis for a Gradient-Based Aerodynamic Shape Optimization Process. Ph.D. Thesis, TU Braunschweig, Braunschweig, Germany, 2021. [Google Scholar]
  210. Viswanathan, V.; Epstein, A.H.; Chiang, Y.M.; Takeuchi, E.; Bradley, M.; Langford, J.; Winter, M. The challenges and opportunities of battery-powered flight. Nature 2022, 601, 519–525. [Google Scholar] [CrossRef]
  211. Lufthansa. The Airbus A380—Lufthansa’s Flagship. Available online: https://www.lufthansagroup.com/en/company/fleet/lufthansa-and-regional-partners/airbus-a380-800.html (accessed on 9 October 2024).
  212. Martins, J.R.R.A.; Ning, A. Engineering Design Optimization; Cambridge University Press: Cambridge, UK, 2022. [Google Scholar] [CrossRef]
  213. Gritzmann, P. Grundlagen der Mathematischen Optimierung: Diskrete Strukturen, Komplexitätstheorie, Konvexitätstheorie, Lineare Optimierung, Simplex-Algorithmus, Dualität; Springer: Fachmedien Wiesbaden, Germany, 2013. [Google Scholar] [CrossRef]
  214. Papageorgiou, M.; Leibold, M.; Buss, M. Optimierung: Statische, dynamische, stochastische Verfahren für die Anwendung; Springer: Berlin/Heidelberg, Germany, 2015. [Google Scholar] [CrossRef]
  215. Stein, O. Grundzüge der Globalen Optimierung; Springer: Berlin/Heidelberg, Germany, 2018. [Google Scholar] [CrossRef]
  216. Stein, O. Grundzüge der Nichtlinearen Optimierung; Springer: Berlin/Heidelberg, Germany, 2018. [Google Scholar] [CrossRef]
  217. Koop, A.; Moock, H. Lineare Optimierung—Eine Anwendungsorientierte Einführung in Operations Research; Springer: Berlin/Heidelberg, Germany, 2018. [Google Scholar] [CrossRef]
  218. Grimme, C.; Bossek, J. Einführung in die Optimierung: Konzepte, Methoden und Anwendungen; Springer: Fachmedien Wiesbaden, Germany, 2018. [Google Scholar] [CrossRef]
  219. Scholz, D. Optimierung Interaktiv: Grundlagen Verstehen, Modelle Erforschen und Verfahren Anwenden; Springer: Berlin/Heidelberg, Germany, 2018. [Google Scholar] [CrossRef]
  220. Jarre, F.; Stoer, J. Optimierung: Einführung in Mathematische Theorie und Methoden; Springer: Berlin/Heidelberg, Germany, 2019. [Google Scholar] [CrossRef]
  221. Dietz, H.M. Mathematik für Wirtschaftswissenschaftler: Band 2: Lineare Algebra und Optimierung; Springer: Berlin/Heidelberg, Germany, 2019. [Google Scholar] [CrossRef]
  222. Schumacher, A. Optimierung Mechanischer Strukturen: Grundlagen und Industrielle Anwendungen; Springer: Berlin/Heidelberg, Germany, 2020. [Google Scholar] [CrossRef]
  223. Xu, B.; Wen, J. Numerical study of spontaneous ignition in pressurized hydrogen release through a length of tube with local contraction. Int. J. Hydrogen Energy 2012, 37, 17571–17579. [Google Scholar] [CrossRef]
  224. Xu, B.; Wen, J. The effect of tube internal geometry on the propensity to spontaneous ignition in pressurized hydrogen release. Int. J. Hydrogen Energy 2014, 39, 20503–20508. [Google Scholar] [CrossRef]
  225. Li, P.; Duan, Q.; Zeng, Q.; Jin, K.; Chen, J.; Sun, J. Experimental study of spontaneous ignition induced by sudden hydrogen release through tubes with different shaped cross-sections. Int. J. Hydrogen Energy 2019, 44, 23821–23831. [Google Scholar] [CrossRef]
  226. Ayi, C.; Sripaul, E.; Vaddiraju, S.; Khan, F. Is hydrogen ignition data from the literature practically observed? Int. J. Hydrogen Energy 2024, 89, 746–759. [Google Scholar] [CrossRef]
  227. Jallais, S.; Vyazmina, E.; Kuznetsov, M.; Grüne, J.; Xu, B.P.; Wen, J.X. Effects of Oxidants on Hydrogen Spontaneous Ignition: Experiments and Modelling. In Proceedings of the International Conference on Hydrogen Safety, Hamburg, Germany, 11–13 September 2017. [Google Scholar]
  228. Landucci, G.; Tugnoli, A.; Cozzani, V. Safety assessment of envisaged systems for automotive hydrogen supply and utilization. Int. J. Hydrogen Energy 2010, 35, 1493–1505. [Google Scholar] [CrossRef]
  229. Moradi, R.; Groth, K.M. Hydrogen storage and delivery: Review of the state of the art technologies and risk and reliability analysis. Int. J. Hydrogen Energy 2019, 44, 12254–12269. [Google Scholar] [CrossRef]
  230. Jaber, M.; Yahya, A.; Arif, A.F.; Jaber, H.; Alkhedher, M. Burst pressure performance comparison of type V hydrogen tanks: Evaluating various shapes and materials. Int. J. Hydrogen Energy 2024, 81, 906–917. [Google Scholar] [CrossRef]
  231. Khzouz, M.; Gkanas, E.I. Hydrogen technologies for mobility and stationary applications: Hydrogen production, storage and infrastructure development. In Renewable Energy-Resources, Challenges and Applications; IntechOpen: London, UK, 2020. [Google Scholar]
  232. Muthukumar, P.; Kumar, A.; Afzal, M.; Bhogilla, S.; Sharma, P.; Parida, A.; Jana, S.; Kumar, E.A.; Pai, R.K.; Jain, I. Review on large-scale hydrogen storage systems for better sustainability. Int. J. Hydrogen Energy 2023, 48, 33223–33259. [Google Scholar] [CrossRef]
  233. Inthavong, K.; Mouritz, A.P.; Dong, J.; Tu, J.Y. Inhalation and deposition of carbon and glass composite fibre in the respiratory airway. J. Aerosol Sci. 2013, 65, 58–68. [Google Scholar] [CrossRef]
  234. Fujitani, Y.; Ikegami, A.; Morikawa, K.; Kumoi, J.; Yano, T.; Watanabe, A.; Shiono, A.; Watanabe, C.; Teramae, N.; Ichihara, G.; et al. Quantitative assessment of nano-plastic aerosol particles emitted during machining of carbon fiber reinforced plastic. J. Hazard. Mater. 2024, 467, 133679. [Google Scholar] [CrossRef]
  235. Airbus. Landing Page for the A350 (Airbus). Available online: https://aircraft.airbus.com/en/aircraft/a350-clean-sheet-clean-start/a350-1000 (accessed on 4 March 2024).
  236. International Standard Atmosphere. Available online: https://www.engineeringtoolbox.com/international-standard-atmosphere-d_985.html (accessed on 7 October 2024).
  237. NASA. U.S. Standard Atmosphere. 1976. Available online: https://ntrs.nasa.gov/citations/19770009539 (accessed on 7 October 2024).
  238. Sicius, H. Noble Gases: Elements of the Eighth Main Group. In Handbook of the Chemical Elements; Springer: Berlin/Heidelberg, Germany, 2024; pp. 449–486. [Google Scholar] [CrossRef]
  239. Sano, Y.; Marty, B.; Burnard, P. Noble Gases in the Atmosphere. In The Noble Gases as Geochemical Tracers; Springer: Berlin/Heidelberg, Germany, 2013; pp. 17–31. [Google Scholar] [CrossRef]
  240. Kaygorodov, M.Y.; Skripnikov, L.V.; Tupitsyn, I.I.; Eliav, E.; Kozhedub, Y.S.; Malyshev, A.V.; Oleynichenko, A.V.; Shabaev, V.M.; Titov, A.V.; Zaitsevskii, A.V. Electron affinity of oganesson. Phys. Rev. 2024, 104, 012819. [Google Scholar] [CrossRef]
  241. Jones, N.; Birch, R. Influence of internal pressure on the impact behavior of steel pipelines. J. Pressure Vessel Technol. 1996, 118, 464–471. [Google Scholar] [CrossRef]
  242. Mitsuishi, H.; Oshino, K.; Watanabe, S. Dynamic crush test on hydrogen pressurized cylinder. Disp 2000, 2005, 1–50. [Google Scholar]
  243. Perfetto, D.; De Luca, A.; Caputo, F.; Lamanna, G. Numerical Investigation on the Impact Resistance of a Hydrogen Fuel Tank. Macromol. Symp. 2022, 404, 2100474. [Google Scholar] [CrossRef]
  244. Farhood, N.H. Low velocity impact simulation of cylindrical section for type IV composite pressure vessels. AIP Conf. Proc. 2021, 2372, 150001. [Google Scholar]
  245. Schwer, L.E.; Holmes, B.S.; Kirkpatrick, S.W. Response and failure of metal tanks from impulsive spot loading: Experiments and calculations. Int. J. Solids Struct. 1988, 24, 817–833. [Google Scholar] [CrossRef]
  246. Sobron, A.; Lundström, D.; Krus, P. A Review of Current Research in Subscale Flight Testing and Analysis of Its Main Practical Challenges. Aerospace 2021, 8, 74. [Google Scholar] [CrossRef]
Figure 3. Gravimetric energy densities of storage systems [63].
Figure 3. Gravimetric energy densities of storage systems [63].
Energies 18 01930 g003
Figure 4. Contrasting hydrogen behavior with an ideal gas using the real gas factor Z [112,114,115,116].
Figure 4. Contrasting hydrogen behavior with an ideal gas using the real gas factor Z [112,114,115,116].
Energies 18 01930 g004
Figure 5. Screenshot of the interactive tool, which compares real and ideal gas behavior and analyzes the influence of pressure and temperature on hydrogen density. The tool is freely available for use at the following link: https://jav-ed.github.io/H2O_Plot/ (accessed on 10 March 2025).
Figure 5. Screenshot of the interactive tool, which compares real and ideal gas behavior and analyzes the influence of pressure and temperature on hydrogen density. The tool is freely available for use at the following link: https://jav-ed.github.io/H2O_Plot/ (accessed on 10 March 2025).
Energies 18 01930 g005
Figure 6. Minimum ignition energy (mJ) of hydrogen in humid air (90% relative humidity) and dry air (0% relative humidity) [151,156,157].
Figure 6. Minimum ignition energy (mJ) of hydrogen in humid air (90% relative humidity) and dry air (0% relative humidity) [151,156,157].
Energies 18 01930 g006
Figure 7. Hydrogen flame at a laminar jet velocity of 47 m / s and a Reynolds number of 837. The image was captured using the Sony DSC D700 with an f/2.4 aperture and no filter [164].
Figure 7. Hydrogen flame at a laminar jet velocity of 47 m / s and a Reynolds number of 837. The image was captured using the Sony DSC D700 with an f/2.4 aperture and no filter [164].
Energies 18 01930 g007
Figure 8. Effect of oxygen content on hydrogen flame visibility. Images show flames at the following different equivalence ratios ϕ : (a) ϕ = 1.0 , (b) ϕ = 0.8 , (c) ϕ = 0.7 , and (d) ϕ = 0.62 . Higher ϕ values indicate closer approximation to stoichiometric combustion. Experimental conditions: jet velocity 33 m / s , Reynolds number 580, turbulent flow, and camera aperture f/2.4 [164].
Figure 8. Effect of oxygen content on hydrogen flame visibility. Images show flames at the following different equivalence ratios ϕ : (a) ϕ = 1.0 , (b) ϕ = 0.8 , (c) ϕ = 0.7 , and (d) ϕ = 0.62 . Higher ϕ values indicate closer approximation to stoichiometric combustion. Experimental conditions: jet velocity 33 m / s , Reynolds number 580, turbulent flow, and camera aperture f/2.4 [164].
Energies 18 01930 g008
Figure 9. Numerical simulation of hydrogen release into the atmosphere from a pressure vessel at 250 bar . Mach number contours are shown at six time points [185].
Figure 9. Numerical simulation of hydrogen release into the atmosphere from a pressure vessel at 250 bar . Mach number contours are shown at six time points [185].
Energies 18 01930 g009
Figure 10. Simulation of hydrogen release into the free environment from a pressure vessel at 250 bar , showing temperature contours at six distinct time points [185].
Figure 10. Simulation of hydrogen release into the free environment from a pressure vessel at 250 bar , showing temperature contours at six distinct time points [185].
Energies 18 01930 g010
Figure 11. Influence of hydrogen container pressure and ambient gas composition on the probability of hydrogen jet self-ignition [227].
Figure 11. Influence of hydrogen container pressure and ambient gas composition on the probability of hydrogen jet self-ignition [227].
Energies 18 01930 g011
Table 2. Constants for approximating the real gas behavior of hydrogen [112].
Table 2. Constants for approximating the real gas behavior of hydrogen [112].
i a i b i c i
1 0.05888460 1.325 1.0
2 0.06136111 1.87 1.0
3 0.002650473 2.5 2.0
4 0.002731125 2.8 2.0
5 0.001802374 2.938 2.42
6 0.001150707 3.14 2.63
7 0.9588528 × 10 4 3.37 3.0
8 0.1109040 × 10 6 3.75 4.0
9 0.1264403 × 10 9 4.0 5.0
Table 3. Examples of factors influencing the choice of filling medium in an experimental and simulative structure test for SWITHs.
Table 3. Examples of factors influencing the choice of filling medium in an experimental and simulative structure test for SWITHs.
Influence on Structural Properties
    1. 
Maximum tolerable bending moment.
    2. 
Maximum elongation.
    3. 
Tensile strength.
    4. 
Torsional stiffness.
    5. 
Young’s modulus.
    6. 
Impact strength.
Impact on Human Health
    1. 
Carcinogenicity.
    2. 
Flammability.
    3. 
Explosiveness.
    4. 
Olfactory effects (impact on nasal mucosa).
    5. 
Radioactivity.
Measurement and Detection Characteristics
    1. 
Visibility.
    2. 
Olfactory detectability.
    2. 
Tactile properties.
    4. 
Responsiveness to electromagnetic fields.
    5. 
Alternative sensory detection methods.
Test Bench Safety Considerations
    1. 
Material compatibility (absence of chemical reactions with test bench components).
Practical Considerations
    1. 
Availability.
    2. 
Environmental impact.
    3. 
Economic efficiency.
Table 4. Combustion types with guide values [80].
Table 4. Combustion types with guide values [80].
Combustion ProcessCombustion RatePressure Rise [bar]
Combustion0.1–50 m / s up to ≈0
Deflagration 0.01 km / s up to ≈10
Explosion 0.1 km / s up to ≈ 10 3
Detonation 1 km / s up to over ≈ 10 5
Table 5. Estimated pressure thresholds for structural damage [80].
Table 5. Estimated pressure thresholds for structural damage [80].
Pressure Rise [bar]Destruction
⪆0.05Window panes
⪆0.1Half-timbered buildings
⪆0.3–0.8Concrete walls, thickness 12–24 cm
Table 6. Conditions for hydrogen to become ignitable to detonable.
Table 6. Conditions for hydrogen to become ignitable to detonable.
  1. 
Presence of sufficient ignition energy or attainment of ignition temperature [148,149,150];
  2. 
Formation of a flammable or detonable gas mixture [144,145,151,152,153];
  3. 
Special reactive cases: presence of chlorine or fluorine [62,63,154,155].
Table 7. Ignition-relevant properties for hydrogen and various other ignitable fluids [62,63].
Table 7. Ignition-relevant properties for hydrogen and various other ignitable fluids [62,63].
SubstanceLower Explosion LimitUpper Explosion LimitFlash PointIgnition TemperatureMinimum Ignition Energy
[Vol% in Air] [Vol% in Air] [°C] [°C] [ mJ ]
Acetylene 1.5 82 136 305 0.019
Ammonia (L)153413265114
Gasoline (L) 0.6 8> 20 240–500 0.8
Natural gas 4.5 13.5 > 188 600 0.3
Carbon monoxide 12.5 75 191 605> 0.3
Methane515 188 595 0.3
Petroleum (L) 0.7 555280 0.25
Propane (L) 2.1 9.5 104 470 0.25
Hydrogen4 75.6 270.8 585 0.017
Table 8. Combustion characteristics and critical values for hydrogen–air mixtures [62,63].
Table 8. Combustion characteristics and critical values for hydrogen–air mixtures [62,63].
Hydrogen–Air Mixture Properties
Lower explosion limit (ignition limit)4 Vol% H2
Lower detonation limit18 Vol% H2
Stoichiometric mixture 29.6 Vol% H2
Upper detonation limit 58.9 Vol% H2
Upper explosion limit (ignition limit) 75.6 Vol% H2
Ignition temperature585 °C ( 858 K )
Minimum ignition energy 0.017 mJ
Maximum laminar flame velocity 3 m / s
Maximum adiabatic combustion temperature≈2200 °C
Table 9. Speed of sound in various media, with gases measured at 20 °C and 1 bar [80].
Table 9. Speed of sound in various media, with gases measured at 20 °C and 1 bar [80].
MediumSpeed of Sound [m/s]
Steel5170
Cast iron3210
Concrete3730
Polymer: Polyvinyl Chloride (PVC)1462
Polymer: Polyethylene (PE)973
Wood4500
Glass5300
Water1437
Diesel1206
Gasoline1062
Air (dry)343
Helium1005
Hydrogen1300
Argon318
Nitrogen349
Methane446
Table 10. Experimental parameters and setups for various studies on hydrogen diffusion ignition [151].
Table 10. Experimental parameters and setups for various studies on hydrogen diffusion ignition [151].
ExperimentBurst Pressure [MPa]Tube Length [mm]Diameter [mm]Cross-Section of Tube
Dryer 2007 [93]1.4–11.338.1–30004/12.7cylinder
Mogi 2008 [197]4–303–3005/10cylinder
Golub 2008 [198]2–1365–18510cylinder/rectangle
Mogi 2009 [199]2–203–5005/10cylinder
Lee 2011 [200]10.8–23.550/100/200/30010.9cylinder
Kim 2013 [190]6.5–11.330010rectangle (visual)
Kim 2013 [201]8–3010–2003cylinder
Kitabayashi 2013 [202]1–8100–420010cylinder
Grune 2014 [203]2.4–24210–11454/5/10cylinder/rectangle (visual)
Duan 2016 [205]2–1180–36015cylinder
Gong 2016 [204]2–936015cylinder
Jiang 2019 [206]5–7220010/15cylinder
Wang 2019 [207]2–12300–300010cylinder
Table 11. Geometrical and environmental factors primarily influencing hydrogen self-ignition, with particular relevance to SWITH design.
Table 11. Geometrical and environmental factors primarily influencing hydrogen self-ignition, with particular relevance to SWITH design.
ParameterOperation
LengthShort
DiameterBig
Compression pressure in the containerLow
Other influencing factors
   
Geometry of the exit area
   
Size of the outer surface into which compressed gaseous hydrogen flows out (can gas accumulations arise?)
   
Gas or medium in the external environment
   
External pressure
   
External temperature
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

Butt, J.A.; Markmiller, J.F.C. Hydrogen Properties and Their Safety Implications for Experimental Testing of Wing Structure-Integrated Hydrogen Tanks. Energies 2025, 18, 1930. https://doi.org/10.3390/en18081930

AMA Style

Butt JA, Markmiller JFC. Hydrogen Properties and Their Safety Implications for Experimental Testing of Wing Structure-Integrated Hydrogen Tanks. Energies. 2025; 18(8):1930. https://doi.org/10.3390/en18081930

Chicago/Turabian Style

Butt, Javed A., and Johannes F. C. Markmiller. 2025. "Hydrogen Properties and Their Safety Implications for Experimental Testing of Wing Structure-Integrated Hydrogen Tanks" Energies 18, no. 8: 1930. https://doi.org/10.3390/en18081930

APA Style

Butt, J. A., & Markmiller, J. F. C. (2025). Hydrogen Properties and Their Safety Implications for Experimental Testing of Wing Structure-Integrated Hydrogen Tanks. Energies, 18(8), 1930. https://doi.org/10.3390/en18081930

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