Compressed hydrogen is a physical storage method for hydrogen gas that involves compressing it to high pressures, typically ranging from 350 to 700 bar (5,000 to 10,000 psi), within specialized high-strength tanks to increase its volumetric energy density for efficient transportation, distribution, and use in energy applications.[1] This approach leverages the gas's inherently high gravimetric energy density—approximately 120 MJ/kg, nearly three times that of gasoline—while enabling rapid refueling times comparable to conventional fuels.[1]The primary applications of compressed hydrogen center on fuel cell electric vehicles (FCEVs), such as cars, buses, and trucks, where it powers proton exchange membrane fuel cells to generate electricity with zero tailpipe emissions, supporting ranges over 300 miles on a single fill of 5 to 13 kg of hydrogen.[1][2] It is also employed in stationary power systems for backup generation and portable devices, with ongoing research focusing on lightweight composite materials like carbon fiber-reinforced polymers for Type IV tanks to enhance durability and reduce weight.[1] The U.S. Department of Energy (DOE) has set technical targets for onboard storage in light-duty vehicles, including a 2025 goal of 1.8 kWh/kg gravimetric capacity (0.055 kg H₂/kg system), 1.3 kWh/L volumetric capacity (0.040 kg H₂/L system), and a cost of $9/kWh (300/kg H₂), though current systems like those in the Toyota Mirai achieve about 5.7 wt% gravimetric density but fall short on volumetric efficiency and affordability.[3][2]Despite its advantages, including fast refueling (under 5 minutes) and compatibility with existing compression infrastructure, compressed hydrogen faces challenges such as low volumetric energy density requiring bulky tanks, high material and manufacturing costs (often exceeding $10/kWh), and safety considerations due to the elevated pressures, which demand robust standards like SAE J2601 for dispensing.[1][2]Compression itself consumes energy—about 4.1 wt% of the hydrogen's content—while pre-cooling for efficient filling adds another 1.8 to 3.6 wt%, contributing to overall system inefficiencies.[2] Historically, compressed hydrogen storage has evolved since the early 2000s with advancements in high-pressure cylinders, reaching 700 bar for vehicular use by companies like Quantum Technologies, enabling practical FCEV deployment while addressing leakage and fatigue issues through improved alloys and composites.[4]
Definition and Overview
Definition
Compressed hydrogen refers to hydrogen gas (H₂) that is stored under high pressure to enhance its volumetric energy density for efficient storage and transportation applications. Typically, this involves pressures ranging from 350 to 700 bar (35 to 70 MPa), achieved using specialized pressure vessels designed to withstand these conditions while ensuring safety and material integrity.[1][5] This method allows hydrogen to remain in its gaseous state at ambient temperatures, distinguishing it from cryogenic storage techniques.Unlike hydrogen stored at ambient pressure, where its low density (approximately 0.09 kg/m³ at standard conditions) limits practical utility, compression significantly boosts the amount of hydrogen that can be contained in a given volume—up to about 40 kg/m³ at 700 bar—without requiring liquefaction. This approach yields system-level gravimetric hydrogen capacities of around 5 wt% in modern composite-overwrapped tanks, enabling energy storage efficiencies suitable for mobile and stationary uses while avoiding the energy-intensive cooling processes associated with liquid hydrogen.[6][7]The commercial application of compressed hydrogen dates back to the early 20th century, when it was first widely used in industrial processes such as oxy-fuel welding and for inflating lighter-than-air balloons, with storage pressures initially around 200 bar in steel cylinders. These early systems laid the groundwork for hydrogen's role as an industrial gas, but it was not until the post-2000 era that advancements in materials like carbon-fiber composites enabled higher pressures (up to 700 bar) and lighter tanks, facilitating its adoption in fuel cell vehicles and emerging clean energy infrastructures.[4][1]The underlying principle of compression relies on the ideal gas law, PV = nRT, where P is pressure, V is volume, n is the number of moles, R is the gas constant, and T is temperature. For density considerations, this can be rearranged to express the mass density \rho = \frac{PM}{RT}, with M as the molar mass of hydrogen, illustrating how increasing P directly enhances \rho and thus the stored energy per unit volume at constant temperature.[6]
Comparison to Other Storage Methods
Compressed hydrogen storage offers a practical alternative to other hydrogen storage methods, such as liquid hydrogen and chemical or metal hydride systems, by operating at ambient temperatures and avoiding complex chemical reactions. Unlike liquid hydrogen, which requires cryogenic cooling to -253°C (20 K) for liquefaction, compressed hydrogen at 700 bar maintains gaseous form without such energy-intensive processes, simplifying handling and reducing boil-off losses during storage. However, this comes at the cost of lower volumetric density: approximately 0.04 kg/L for compressed hydrogen systems at 700 bar and 25°C, compared to ~0.04 kg/L for liquid hydrogen systems.[8][9] This density similarity means both occupy comparable volumes for equivalent hydrogen mass in practical systems, though compressed eliminates the need for specialized cryogenic infrastructure.[6]In contrast to chemical or metal hydride storage, compressed hydrogen provides a fully reversible process without requiring heat input for release or regeneration, making refueling as straightforward as repressurization. Metal hydrides, such as magnesium-based or complex hydrides like sodium alanate, can achieve higher volumetric densities (up to 0.05-0.10 kg/L or more) at near-ambient conditions and lower pressures, potentially reducing vessel requirements. However, hydrides often suffer from slower kinetics, the need for thermal management during absorption and desorption, and in some cases irreversibility, complicating onboard reversibility compared to the simpler, robust high-pressure vessels used in compressed systems.[10][11]System-level energy densities further highlight these trade-offs. Current 700 bar compressed hydrogen systems deliver about 0.9 kWh/L and 1.4 kWh/kg as of 2025, with targets of 1.3 kWh/L and 1.8 kWh/kg, reflecting the impact of tank weight and volume, which is significantly lower than gasoline's 8.9 kWh/L but competitive for hydrogen's high gravimetric potential. Liquid hydrogen systems achieve about 1.3 kWh/L, higher than current compressed but challenged by insulation, while advanced hydrides aim for 1.5-2.0 kWh/L but remain challenged by material costs and cycle life. As of 2025, compressed system costs are projected at $12-15/kWh, with gravimetric efficiencies reaching 5-6 wt% in vehicles like the Toyota Mirai.[12][1][13]
These metrics underscore compressed hydrogen's suitability for short-term, mobile applications like fuel cell vehicles, where quick refueling and simplicity outweigh volumetric limitations, but it proves less efficient for long-term stationary storage compared to hydrides or underground options that minimize compression energy over time.[11][1]
Properties
Physical Properties
Compressed hydrogen exhibits significant variations in density with pressure, behaving as a real gas where density increases nearly linearly but with deviations due to compressibility factors. At standard temperature and pressure (STP), hydrogen density is approximately 0.0899 kg/m³, but under compression to 700 bar at 15°C, it reaches about 40 kg/m³, enabling more efficient volumetric storage compared to atmospheric conditions.[14][15] This increase follows the equation of state for real gases, such as the Redlich-Kwong model, accounting for intermolecular forces at high pressures.[16]The viscosity of compressed hydrogen remains low, facilitating flow in storage and distribution systems. At high pressures around 700bar and 15°C, dynamic viscosity is approximately 10 μPa·s, slightly higher than the 8.8 μPa·s at atmospheric pressure due to enhanced molecular interactions.[17][18] Thermal conductivity is also notable, valued at about 0.17 W/m·K near ambient conditions and high pressure, which supports effective heat transfer during compression processes.[19]In terms of phase behavior, compressed hydrogen remains in the gaseous state well above its critical point of 33.2 K (-239.95°C) and 12.93 bar, where the distinction between liquid and gas phases vanishes. At typical storage temperatures around 15–20°C and pressures up to 700 bar, it enters the supercritical regime, exhibiting non-ideal behaviors such as the Joule-Thomson effect, which can lead to temperature changes during pressure reductions in handling.[20][16] The specific heat capacity at constant pressure (Cp) is approximately 14.3 J/g·K (or 28.6 J/mol·K) at STP, increasing modestly under compression to around 14.5 J/g·K at 700 bar due to real gas effects.[21][22]
Pressure (bar)
Density (kg/m³) at 15°C
Viscosity (μPa·s) at 15°C
Cp (J/g·K)
1
0.084
8.8
14.3
350
24
9.5
14.4
700
40
10
14.5
These properties contrast with liquid hydrogen, which achieves higher densities (~70 kg/m³) but requires cryogenic conditions.[6]
Thermodynamic Properties
Compressed hydrogen exhibits significant deviations from ideal gas behavior at high pressures, primarily characterized by the compressibility factor Z = \frac{PV}{RT}, which quantifies the departure from ideality. At 700 bar and approximately 300 K, Z \approx 1.4, indicating that the actual volume occupied by hydrogen is larger than predicted by the ideal gas law due to intermolecular repulsive forces dominating over attractive ones.[23] This value is derived from experimental data and equation-of-state models, such as the one provided by the National Institute of Standards and Technology (NIST). The van der Waals equation of state approximates this behavior for hydrogen:\left( P + \frac{a}{V^2} \right) (V - b) = RT,where a = 0.247 \, \mathrm{L^2 \cdot bar / mol^2} and b = 0.0266 \, \mathrm{L/mol} are the van der Waals constants specific to H_2, accounting for molecular volume exclusion and weak attractive interactions.[24]In compression processes, the thermodynamic work required varies between adiabatic and isothermal conditions, impacting overall energy efficiency. For reversible adiabatic compression of diatomic hydrogen (\gamma \approx 1.41), the work per mole is given byW = \frac{\gamma}{\gamma - 1} RT_1 \left[ \left( \frac{P_2}{P_1} \right)^{(\gamma - 1)/\gamma} - 1 \right],where \gamma = C_p / C_v reflects the heat capacity ratio. Practical multi-stage compression with intercooling requires more energy than the ideal isothermal process due to mechanical inefficiencies and real-gas effects, typically consuming 6–8% of the hydrogen's energy content.[25] This difference arises from the entropy increase in irreversible paths compared to the constant-temperature isothermal case.[26]The heat generated during compression necessitates intercooling in multi-stage setups to limit temperature rises to about 200°C, preventing excessive energy losses and material stress. Under high pressure, changes in thermodynamic properties include shifts in enthalpy and entropy; for instance, reflecting real-gas corrections beyond ideal behavior, while entropy decreases due to reduced volume availability, \Delta S \approx -R \ln(P_2 / P_1) adjusted for non-ideality. These properties are critical for evaluating storage efficiency and are tabulated in NIST reference data for precise engineering applications.[27][16]
Compression and Storage
Compression Processes
Compressed hydrogen is typically produced through mechanical compression processes that elevate the gas from low inlet pressures to high storage pressures, often ranging from 200 to 700 bar for vehicular and industrial applications. The primary technologies employed are multi-stage reciprocating compressors, which dominate due to their reliability and capacity for high-pressure delivery. These systems divide the compression into several stages to manage temperature rise and improve efficiency, with intercooling between stages to approximate isothermal conditions and reduce overall work input.[28]Multi-stage reciprocating compressors are the standard for achieving pressures of 200–700 bar, utilizing piston-cylinder mechanisms driven by electric motors or turbines. For instance, four to five stages are commonly used to reach 700 bar, with each stage featuring a pressure ratio of about 2:1 to minimize energy losses from heat generation. Intercooling stages cool the gas back toward ambient temperature, enhancing isentropic efficiency to approximately 70–80% in well-designed systems.[28][29]For applications requiring ultra-high purity, such as fuel cell production or refueling stations, oil-free alternatives like diaphragm and ionic liquid compressors are preferred to avoid contamination from lubricants. Diaphragm compressors use a flexible metal diaphragm separated from hydraulic oil to compress the gas, achieving up to 1000 bar while maintaining high hydrogen purity. Ionic liquid compressors employ a non-volatile ionic liquid as a piston substitute, enabling oil-free operation and pressures up to 1000 bar with minimal gas solubility and contamination risks.[29][30]The energy required for compressing hydrogen to 700 bar is approximately 3–4 kWh per kg of H₂, accounting for 10–15% of the hydrogen's lower heating value of 120 MJ/kg. This consumption arises from the thermodynamic work of compression, as outlined in related properties, with practical efficiencies influenced by staging and cooling.[28]A typical compressionprocess begins with intake of hydrogen gas at 10–20 bar from production sources like electrolyzers, followed by progressive staging in reciprocating or oil-free units. Each stage increases pressure incrementally, with intercoolers reducing gas temperature between stages and aftercoolers bringing the final output to near-ambient conditions before storage, ensuring safe and efficient handling.[28]
Storage Systems and Materials
Compressed hydrogen storage relies on high-pressure vessels designed to contain the gas safely while optimizing weight, volume, and cost for applications such as vehicular and stationary systems. These vessels are classified into four main types (I through IV) based on their construction materials and reinforcement, with each type balancing pressure capacity, durability, and gravimetric efficiency. Type I and II vessels use metallic materials for simplicity and lower cost, while Types III and IV incorporate composites for higher pressures and reduced weight.[31]Type I vessels are fully metallic, typically constructed from steel or aluminum, and are suited for lower operating pressures of 200–350 bar. These all-metal cylinders provide cost-effective storage but are significantly heavier due to the thick walls required to withstand pressure, resulting in gravimetric hydrogen densities of approximately 0.5–1 kg of H₂ per 100 kg of vessel weight. Type II vessels feature a metallic liner (often aluminum) with partial composite hoop wrapping around the cylindrical section, offering modest weight savings over Type I while maintaining compatibility with 200–350 bar pressures; they remain cost-competitive for non-mobile uses but exhibit similar low gravimetric densities.[7][32]Type IV vessels represent the advanced standard for high-pressure storage, consisting of a non-metallic plastic liner (typically high-density polyethylene) overwrapped with carbon fiber-reinforced polymer (CFRP) composites. Rated for up to 700 bar, these cylinders achieve substantial weight reductions—approximately 20% of the mass of equivalent steel Type I vessels—enabling higher gravimetric efficiencies around 4–5 wt% H₂, which is critical for mobile applications like fuel cell vehicles. The polymer liner prevents hydrogen permeation, while the carbon fiber wrap bears the primary load, enhancing burst resistance and fatigue life.[33][13]Certification of these vessels ensures safety through rigorous testing under standards like ISO 19881 for gaseous hydrogen land vehicle fuel containers, which mandates minimum burst pressures of 2.25–3.5 times the operating pressure depending on construction type (e.g., 2.25 times for non-welded composite types and up to 3.5 times for welded metallic types). This includes hydrostatic pressure tests, cyclic loading to simulate refueling, and environmental exposure assessments to verify integrity under real-world conditions. Compliance with ISO 19881, alongside regional codes like SAE J2579, is required for automotive deployment.[34]Post-2020 innovations have focused on all-composite designs, such as Type V vessels that are linerless and fully constructed from carbon fiber-reinforced polymers, enabling operation at 1000 bar to boost volumetric density. These Type V tanks improve gravimetric capacity to 6–7 wt% H₂ by optimizing fiber angles and resin systems, surpassing traditional 700 bar limits while maintaining safety margins; prototypes from companies like Infinite Composites demonstrate feasibility for next-generation storage in heavy-duty transport and space applications as of 2025.[35][3]
Applications
Transportation Uses
Compressed hydrogen is widely utilized in fuel cell electric vehicles (FCEVs), where it is stored at pressures up to 700 bar to power onboard fuel cells that generate electricity for electric motors. Passenger cars like the Toyota Mirai exemplify this application, with the second-generation model launched in 2021 featuring three carbon-fiber-reinforced tanks holding approximately 5.6 kg of hydrogen, enabling a driving range of approximately 650 km (EPA-estimated 402 miles) depending on conditions.[36] The 2025 Mirai update maintains this storage capacity and range of up to 647 km while incorporating refinements for improved efficiency and standard features.[37] These vehicles achieve zero tailpipe emissions, producing only water vapor, making them suitable for urban and highway driving in regions with supporting refueling infrastructure.In heavy-duty transportation, compressed hydrogen supports zero-emission operations for trucks and buses, often stored at 350–700 bar to accommodate larger payloads and longer routes. The Hyundai XCIENTFuel Cell truck, first deployed commercially in Europe in 2020 and expanded to North America in 2025, uses ten high-pressure tanks to store up to 70 kg of hydrogen, providing a range of approximately 720 km even when fully loaded.[38] For buses, similar systems enable fleet operations, with roof-mounted tanks typically at 350 bar sufficient for daily routes of 300–500 km.[39] These applications leverage compressed hydrogen's high energy density by weight, allowing heavy-duty vehicles to maintain payload capacities comparable to diesel counterparts while reducing greenhouse gas emissions in long-haul and public transit sectors.[40]Refueling with compressed hydrogen at 700 bar typically takes 3–5 minutes to fill a light-duty FCEV tank, comparable to conventional gasoline refueling and significantly faster than the 30–60 minutes required for fast-charging battery electric vehicles to achieve similar energy levels.[41] This rapid turnaround supports high-utilization scenarios in transportation. Globally, the FCEV fleet exceeded 100,000 vehicles by the end of 2024, with around 5,000 new car sales that year, driven by mandates such as California's Advanced Clean Cars II regulation aiming for 35% zero-emission vehicle sales by 2026 and the European Union's CO2 emission standards promoting hydrogen adoption.[42]
Stationary and Industrial Uses
Compressed hydrogen serves as a reliable energy storage medium in stationary backup power systems, particularly for critical infrastructure like data centers. These systems utilize high-pressure storage at 350 bar to enable uninterruptible power supplies through fuel cells, offering autonomy durations of 24 to 72 hours during outages.[43][1] Such setups replace traditional diesel generators, providing clean, on-site power without emissions, with storage tanks sized to hold tens of tonnes of hydrogen for modular data center capacities up to 20–30 MW.[43]In industrial applications, compressed hydrogen acts as a key feedstock for processes like ammonia synthesis via the Haber-Bosch method and petroleum refining. For ammonia production, hydrogen is compressed on-site from typical electrolyzer output pressures of around 30 bar to 100–300 bar to facilitate the high-pressure reaction with nitrogen over iron catalysts at 400–500°C.[44] In oil refining, particularly hydrotreating and hydrocracking, hydrogen is compressed to 20–200 bar to remove sulfur and upgrade heavy feedstocks, enhancing product yields and meeting environmental standards.[45] These on-site compression stages ensure efficient integration of hydrogen into existing plant infrastructure, supporting large-scale chemical and fuel production.[44]Compressed hydrogen storage at 100–200 bar plays a vital role in grid balancing by integrating renewable energy sources, storing excess electricity-generated hydrogen for later release via gas turbines or fuel cells. This approach mitigates intermittency from wind and solar, enabling seasonal energy shifting in power systems. In Germany, projects like Bosch's 2.5 MW electrolyzer-storage hybrid, operational since November 2025, exemplify this by combining on-site production and compression for grid support in industrial regions.[46][47]The round-trip efficiency for stationary compressed hydrogen systems, encompassing compression, storage, and decompression cycles, typically ranges from 40% to 50%, reflecting losses in electrolysis and fuel cell conversion but providing long-duration storage advantages over batteries.[48]
Safety and Handling
Associated Hazards
Compressed hydrogen poses significant hazards due to its high storage pressures, typically up to 700 bar, which can lead to catastrophic vessel failure if integrity is compromised. A rupture of a 700 bar tank can release hydrogen rapidly, generating a blast wave with overpressures capable of causing structural damage and injury over considerable distances; for instance, simulations of a 60-liter onboard tank rupture indicate a no-harm overpressure radius (1.35 kPa) of approximately 75-82 meters, while injury thresholds (higher overpressures) extend to about 13 meters.[49] Such failures may result in jet fires or explosions if the released hydrogen ignites, amplifying the risk of thermal radiation and fragmentation hazards.[49]Hydrogen embrittlement represents another critical risk, where atomic hydrogen diffuses into metallic components, particularly high-strength steels used in storage vessels, leading to reduced ductility and crack propagation. This process involves hydrogen atoms entering the metal lattice, recombining into molecular hydrogen at defects like inclusions or carbides, which generates internal pressure and initiates microcracks; in high-strength steels, even low hydrogen concentrations (0.001-0.1 ppm) can cause slow crack growth under tensile stress, culminating in brittle fracture.[50] The phenomenon is exacerbated in environments with sustained hydrogen exposure, affecting material toughness and potentially leading to sudden failures in pressure-containing systems.[50]Leakage from compressed hydrogen systems heightens ignition risks owing to hydrogen's broad flammability range of 4-75% by volume in air and its extremely low minimum ignition energy of 0.02 mJ, allowing ignition from minor sparks or static discharge.[51] Additionally, hydrogen flames are nearly invisible in daylight due to low luminosity and minimal radiant heat emission, complicating detection and increasing the likelihood of unnoticed exposure or escalation to explosion.[52] These properties make even small leaks potentially hazardous, as hydrogen disperses rapidly but can accumulate in confined spaces to form explosive mixtures.A notable incident illustrating these hazards occurred in June 2019 at the Kjørbo hydrogen refueling station near Oslo, Norway, where a hydrogen leak from a high-pressure storage tank, caused by an assembly error in a tank plug, led to an explosion that injured three people via a pressure wave strong enough to deploy vehicle airbags.[53] The event underscores the vulnerability of high-pressure systems to human error, resulting in over $4 million in damages and temporary shutdowns of similar stations.[54]
Standards and Protocols
The ISO 19880 series establishes international standards for gaseous hydrogen fueling stations, with a focus on safety and performance for compressed hydrogen dispensers used in land vehicle refueling. This series includes specifications for H70 nozzles designed for 700 bar (70 MPa) fueling, incorporating features such as breakaway mechanisms and thermal flow compensation to ensure secure connections and prevent leaks during high-pressure transfers. Flow rates are regulated to support efficient refueling, typically up to 3.6 kg/min for standard operations and higher for heavy-duty applications, balancing speed with safety to avoid overheating or overpressurization.UN ECE Regulation 134 outlines requirements for the type approval of compressed hydrogen storage systems in vehicles, emphasizing rigorous testing to verify structural integrity under operational and extreme conditions. Tanks must undergo proof pressure testing at 1.25 times the nominal working pressure (NWP), followed by burst testing at a minimum of 225% NWP to confirm they exceed failure thresholds without rupture. Additionally, crash simulations are mandated, including frontal, side, and rear impact tests, followed by post-crash leak assessments to ensure no hydrogen release exceeds 3.24% of the initial stored mass over 90 minutes, thereby mitigating risks during accidents.[55]Leak detection protocols for compressed hydrogen facilities rely on specialized sensors capable of detecting hydrogen concentrations at parts per million (ppm) sensitivity, typically in the 1,000 to 4,000 ppm range for early warning of potential leaks before reaching flammable levels (4% lower explosive limit). These sensors must be positioned to monitor high-risk areas such as storage vessels and piping, often integrated with automatic shutdown systems. Ventilation requirements, as specified in the NFPA 2 Hydrogen Technologies Code, mandate mechanical exhaust systems providing at least 1 cubic foot per minute per square foot (0.005 m³/s per m²) of floor area in enclosed spaces to dilute and disperse any released hydrogen, preventing accumulation and ignition.[56]Operator training and certification programs from the CSA Group address safe handling of compressed hydrogen, with a strong emphasis on standardized protocols to minimize human error. These include detailed purge procedures using inert gases like nitrogen to displace residual air from systems before introducing hydrogen, thereby avoiding air admixture that could form explosive mixtures (e.g., 4-75% hydrogen in air). Certification requires demonstrating proficiency in these procedures, along with emergency response and equipment inspection, ensuring compliance with codes like CAN/BNQ 1784-000, the Canadian Hydrogen Installation Code.[57][58]
Infrastructure
Production Integration
Compressed hydrogen production is predominantly linked to steam methane reforming (SMR), which accounts for approximately 76% of global hydrogen output.[59] In the SMR process, hydrogen is generated from natural gas and steam at pressures typically ranging from 20 to 40 bar, followed by purification through pressure swing adsorption (PSA) to achieve high purity before further compression for storage and distribution.[60][61]For renewable pathways, integration with proton exchange membrane (PEM) electrolyzers is increasingly common, where low-pressure hydrogen (around 30 bar) is produced from water using renewable electricity and then compressed to 350–700 bar using co-located compressor systems for efficient storage.[62][1] This setup minimizes energy losses and supports seamless scaling in green hydrogen facilities.Emerging trends in green hydrogen emphasize compressed storage for export, with low-emissions production on track to reach 1 Mt by late 2025, of which 60% derives from electrolysis.[63][64]The economic integration of compression adds approximately $0.1–0.3/kg to overall production costs for green hydrogen, which range from $3–6/kg as of 2025 depending on renewable energy prices and scale, underscoring the need for efficiency improvements in compressor technologies.[65][66]
Distribution and Refueling
Compressed hydrogen is primarily distributed via tube trailers for short-haul deliveries, consisting of cascades of cylinders pressurized to 200–350 bar, with typical capacities of 300–500 kg per trailer.[67] These trailers are well-suited for regional transport to refueling stations or industrial sites, leveraging existing roadinfrastructure while adhering to safety regulations for high-pressure gas handling.[68]Pipeline transport of pure compressed hydrogen remains rare due to the risk of hydrogen embrittlement, which can degrade steel pipelines and increase fracture susceptibility.[69] However, emerging applications involve blending hydrogen into natural gas pipelines at concentrations up to 20% by volume, supported by compression boosters to maintain flow and pressure along the network.[70]Refueling infrastructure for compressed hydrogen features high-pressure stations equipped with 700 bar dispensers, often incorporating pre-cooling systems to -40°C to enable rapid fills while controlling temperature rise in vehicle tanks.[71] As of late 2025, over 1,300 such stations operate globally (with recent closures in Europe highlighting demand challenges), with concentrations in Japan (around 160 stations), Germany (about 90), and California (around 50 operational).[72][73][74]The cascade filling method is commonly employed at these stations, involving sequential pressure equalization from multiple high-pressure storage banks to the vehicle's tank, which achieves up to 90% fill efficiency by minimizing energy loss and optimizing mass transfer.[75] This approach enhances overall station throughput and supports compliance with international refueling protocols.[76]