Abstract

This review explores the advancements in solar technologies, encompassing production methods, storage systems, and their integration with renewable energy solutions. It examines the primary hydrogen production approaches, including thermochemical, photochemical, and biological methods. Thermochemical methods, though highly efficient, require advanced materials and complex reactor designs, while photochemical methods offer a simpler alternative but suffer from low conversion efficiencies. Biological hydrogen production presents a low-cost option but faces limitations in scalability and production rates. The review also highlights innovative hydrogen storage technologies, such as metal hydrides, metal-organic frameworks, and liquid organic hydrogen carriers, which address the intermittency of solar energy and offer scalable storage solutions. Additionally, the potential of hybrid energy systems that integrate solar hydrogen with photovoltaics, thermal energy systems, battery storage, and smart grids is emphasized. Despite technical and economic barriers, ongoing advancements in catalyst development, material optimization, and artificial intelligence-driven energy management systems are accelerating the adoption of solar hydrogen technologies. These innovations position solar hydrogen as a pivotal solution for achieving a sustainable and low-carbon energy future.

1. Introduction

The global energy landscape is currently undergoing a significant transformation due to the increasing demand for sustainable and environmentally friendly energy solutions, which is driven by several factors [1], including the growing global population, rapid industrialization, and the urgent need to mitigate the effects of climate change. Traditional energy sources, such as fossil fuels, contribute heavily to greenhouse gas emissions and are limited in supply [2], making them unsustainable in the long term. Wind, solar, and hydropower offer promising alternatives that can significantly reduce the environmental impact of energy production, in which solar energy stands out due to its abundance and geographical flexibility, which can be captured in almost any location on Earth [3], making it a flexible and widely usable option for energy generation. However, solar energy is not always available because it depends on daylight and weather conditions, which creates a significant challenge, requiring the development of efficient energy storage systems to ensure a consistent and reliable energy supply. Solar energy can be captured and converted into various forms, including electrical energy via photovoltaics (PVs), thermal energy through solar heating systems, and chemical energy in the form of solar fuels, in which the conversion of solar energy into chemical energy represents a promising strategy for overcoming the challenges of intermittency and storage [4]. Solar fuels, such as hydrogen, store solar energy in chemical bonds that can be released on demand, providing a flexible and long-term energy storage solution. As a clean energy carrier, hydrogen can be used in fuel cells to produce electricity with water as the only byproduct, making it an attractive option for reducing carbon emissions in sectors such as transportation, industry, and power generation [5]. Furthermore, hydrogen can be stored in compressed, liquefied, or chemically bonded forms, providing a versatile means of energy storage and transport. One of the most promising avenues for producing hydrogen sustainably is through solar hydrogen production, which directly or indirectly uses solar energy to split water into hydrogen and oxygen. Unlike traditional hydrogen production methods that depend on fossil fuels and produce significant carbon emissions, solar hydrogen production is environmentally friendly and uses an unlimited energy source [6]. Solar hydrogen production can be achieved through several processes, including thermochemical water splitting, photochemical reactions, and biological processes. In addition, hydrogen can serve as both a fuel and an energy storage medium, and its ability to be stored for long periods enables it to bridge the gap between solar energy availability and demand, effectively addressing the intermittency of solar power [7], which makes it a valuable resource not only in energy production but also in stabilizing energy grids and providing backup power in off-grid applications. Despite the significant promise of solar hydrogen, there are still numerous technical and economic barriers that must be overcome before it can be deployed at scale. Challenges [8] such as low conversion efficiencies, high production costs, and the need for advanced materials in catalytic processes are currently limiting the widespread adoption of solar hydrogen technologies. However, ongoing research in areas such as catalyst development [9, 10], reactor design [11, 12], and integration with renewable energy systems [13–15] is steadily improving the viability of these technologies. This review will provide a comprehensive overview of the current state of solar hydrogen production, storage technologies, and systems integration, with a focus on the major approaches including thermochemical, photochemical, and biological methods as illustrated in Fig. 1, which presents a graphical abstract of solar-powered hydrogen technologies. By examining the underlying principles, recent advancements, and technical challenges associated with each method, the aim of this study is to highlight the potential of solar hydrogen as a key component in the future of renewable energy storage.

A diagram of solar-powered hydrogen technology covering three areas:Production: Includes thermochemical, photochemical, and biological methods focusing on efficiency and process improvements. Storage: Highlights methods like metal hydrides, MOFs, and LOHCs, with a focus on thermodynamics and kinetics and their applications. Energy Integration: Explores PV integration, battery synergies, smart grids, and case studies like SOLAR-JET and FH2R.
Figure 1.

Graphical abstract of solar-powered hydrogen technologies.

2. Solar hydrogen production technologies

Solar energy can be converted into hydrogen through three primary methods (as shown in Fig. 2): thermochemical, photochemical, and biological processes. Thermochemical production involves high-temperature reactions, often using metal oxides, to split water into hydrogen and oxygen, typically driven by concentrated solar power (CSP). Photochemical methods use solar energy to directly drive water-splitting reactions via photocatalysts, offering a potentially simpler and lower-cost route to hydrogen. Lastly, biological processes utilize microorganisms, such as algae or bacteria, to produce hydrogen from sunlight through their natural metabolic activities.

 A diagram illustrating hydrogen production methods, divided into three categories: Thermochemical: Includes high-temperature reactions, metal oxide cycles, sulfur-iodine cycles, and hybrid cycles, with challenges like material degradation and heat transfer. Photochemical: Focuses on light-driven water splitting, using single-component catalysts, Type-II heterojunctions, and Z-scheme systems, with challenges such as low efficiency and catalyst stability. Biological: Involves microbial processes using hydrogenase and nitrogenase enzymes, highlighting challenges like low production rates and advancements through genetic and metabolic engineering.
Figure 2.

An illustration of hydrogen production methods.

2.1 Thermochemical hydrogen production

Thermochemical hydrogen production is a promising method for typically utilizing CSP to drive high-temperature chemical reactions, splitting water into hydrogen and oxygen, which utilizes solar energy to heat reactants to temperatures that facilitate water decomposition, typically involving such as sulfur-iodine (S-I) [16, 17], hybrid sulfur (HyS) [18, 19], calcium-bromine [20], and metal oxide reduction-oxidation (redox) cycles [21–27]. Among these, metal oxide cycles have gained significant attention due to their high efficiency, scalability, and relative simplicity [28]. These cycles, such as ZnO/Zn, CeO2/Ce2O3, and Fe3O4/FeO, operate through a straightforward two-step redox mechanism that effectively converts solar thermal energy into chemical energy [21]. Unlike the S-I and HyS cycles, which involve corrosive chemicals and complex multi-step reactions, metal oxide cycles rely on widely available and cost-effective materials like ZnO and CeO2, making them practical for industrial-scale applications [24]. In comparison, methods like S-I and HyS, while capable of high yields, involve greater operational complexity and higher costs, limiting their scalability [17]. Metal oxide cycles thus strike an ideal balance between efficiency, simplicity, and scalability, positioning them as a leading candidate for sustainable and large-scale solar hydrogen production. Table 1 provides a comprehensive overview of thermochemical hydrogen production methods, summarizing their efficiency, production rates, material durability, potential advantages, and associated challenges.

Table 1.

Summary of thermochemical hydrogen production methods, efficiency, production rate, material durability, and challenges.

MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
Metal Oxide Redox Cycles (ZnO/Zn, CeO2/Ce2O3, Fe3O4/FeO)30–5050–1001000 cyclesHigh efficiency, well-studiedMaterial degradation, high temperatures[21–27]
S-I Cycle35–4540–90500–1000 cyclesProven cycle, driven by CSPCorrosive chemicals, expensive materials[16, 17]
HyS Cycle40–5060–110500–1000 cyclesCombines electrochemical and thermochemicalRequires both heat and electricity[18, 19]
Ca-Br Cycle25–3530–70500–1000 cyclesLower temperature processLower efficiency, complex handling[20]
Catalytic Coatings on Metal Oxides30–5050–1001000 cyclesImproves reaction rates, reduces temperaturesHigh cost of precious metals[29, 30]
Optimizing Particle Size30–5050–1001000 cyclesIncreases surface area, enhances reactionsScaling up for industrial use[31]
Reactor Configurations (Fluidized Bed, Rotating Particle)35–5560–1201500 cyclesImproves heat and mass transferComplex engineering for scaling[32]
High-Performance Materials (Ceramics and Refractory Metals)30–5050–1001000 cyclesHigh thermal resistance, Withstands extreme temperaturesExpensive, prone to failure under cycling, oxidation concerns[33, 34]
Coating Technologies to Protect Metal Oxides30–5050–1001200 cyclesPrevents degradation and sinteringRequires robust materials for high temps[35]
Parabolic Troughs35–5060–1101500 cyclesEfficient solar concentrationRequires large solar field[36]
Central Tower Receivers40–6070–1302000 cyclesMaximizes solar concentrationRequires consistent sunlight[37]
Recuperative Heat Exchangers30–5050–1001200 cyclesImproves energy efficiencyDesign complexity[38]
Membrane-Based Separation Systems (Ceramic Membranes, Perovskite Oxides)30–5050–1001500 cyclesEfficient hydrogen separationHigh manufacturing cost[39–41]
Rotating Particle Reactors and Fluidized Bed35–5560–1201500 cyclesOptimizes reaction conditionsComplex integration and design[42, 43]
Nanostructured Metal Oxides and Composite Materials35–5560–1201500 cyclesEnhances catalytic propertiesExpensive to scale up[22, 23, 27]
MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
Metal Oxide Redox Cycles (ZnO/Zn, CeO2/Ce2O3, Fe3O4/FeO)30–5050–1001000 cyclesHigh efficiency, well-studiedMaterial degradation, high temperatures[21–27]
S-I Cycle35–4540–90500–1000 cyclesProven cycle, driven by CSPCorrosive chemicals, expensive materials[16, 17]
HyS Cycle40–5060–110500–1000 cyclesCombines electrochemical and thermochemicalRequires both heat and electricity[18, 19]
Ca-Br Cycle25–3530–70500–1000 cyclesLower temperature processLower efficiency, complex handling[20]
Catalytic Coatings on Metal Oxides30–5050–1001000 cyclesImproves reaction rates, reduces temperaturesHigh cost of precious metals[29, 30]
Optimizing Particle Size30–5050–1001000 cyclesIncreases surface area, enhances reactionsScaling up for industrial use[31]
Reactor Configurations (Fluidized Bed, Rotating Particle)35–5560–1201500 cyclesImproves heat and mass transferComplex engineering for scaling[32]
High-Performance Materials (Ceramics and Refractory Metals)30–5050–1001000 cyclesHigh thermal resistance, Withstands extreme temperaturesExpensive, prone to failure under cycling, oxidation concerns[33, 34]
Coating Technologies to Protect Metal Oxides30–5050–1001200 cyclesPrevents degradation and sinteringRequires robust materials for high temps[35]
Parabolic Troughs35–5060–1101500 cyclesEfficient solar concentrationRequires large solar field[36]
Central Tower Receivers40–6070–1302000 cyclesMaximizes solar concentrationRequires consistent sunlight[37]
Recuperative Heat Exchangers30–5050–1001200 cyclesImproves energy efficiencyDesign complexity[38]
Membrane-Based Separation Systems (Ceramic Membranes, Perovskite Oxides)30–5050–1001500 cyclesEfficient hydrogen separationHigh manufacturing cost[39–41]
Rotating Particle Reactors and Fluidized Bed35–5560–1201500 cyclesOptimizes reaction conditionsComplex integration and design[42, 43]
Nanostructured Metal Oxides and Composite Materials35–5560–1201500 cyclesEnhances catalytic propertiesExpensive to scale up[22, 23, 27]
Table 1.

Summary of thermochemical hydrogen production methods, efficiency, production rate, material durability, and challenges.

MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
Metal Oxide Redox Cycles (ZnO/Zn, CeO2/Ce2O3, Fe3O4/FeO)30–5050–1001000 cyclesHigh efficiency, well-studiedMaterial degradation, high temperatures[21–27]
S-I Cycle35–4540–90500–1000 cyclesProven cycle, driven by CSPCorrosive chemicals, expensive materials[16, 17]
HyS Cycle40–5060–110500–1000 cyclesCombines electrochemical and thermochemicalRequires both heat and electricity[18, 19]
Ca-Br Cycle25–3530–70500–1000 cyclesLower temperature processLower efficiency, complex handling[20]
Catalytic Coatings on Metal Oxides30–5050–1001000 cyclesImproves reaction rates, reduces temperaturesHigh cost of precious metals[29, 30]
Optimizing Particle Size30–5050–1001000 cyclesIncreases surface area, enhances reactionsScaling up for industrial use[31]
Reactor Configurations (Fluidized Bed, Rotating Particle)35–5560–1201500 cyclesImproves heat and mass transferComplex engineering for scaling[32]
High-Performance Materials (Ceramics and Refractory Metals)30–5050–1001000 cyclesHigh thermal resistance, Withstands extreme temperaturesExpensive, prone to failure under cycling, oxidation concerns[33, 34]
Coating Technologies to Protect Metal Oxides30–5050–1001200 cyclesPrevents degradation and sinteringRequires robust materials for high temps[35]
Parabolic Troughs35–5060–1101500 cyclesEfficient solar concentrationRequires large solar field[36]
Central Tower Receivers40–6070–1302000 cyclesMaximizes solar concentrationRequires consistent sunlight[37]
Recuperative Heat Exchangers30–5050–1001200 cyclesImproves energy efficiencyDesign complexity[38]
Membrane-Based Separation Systems (Ceramic Membranes, Perovskite Oxides)30–5050–1001500 cyclesEfficient hydrogen separationHigh manufacturing cost[39–41]
Rotating Particle Reactors and Fluidized Bed35–5560–1201500 cyclesOptimizes reaction conditionsComplex integration and design[42, 43]
Nanostructured Metal Oxides and Composite Materials35–5560–1201500 cyclesEnhances catalytic propertiesExpensive to scale up[22, 23, 27]
MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
Metal Oxide Redox Cycles (ZnO/Zn, CeO2/Ce2O3, Fe3O4/FeO)30–5050–1001000 cyclesHigh efficiency, well-studiedMaterial degradation, high temperatures[21–27]
S-I Cycle35–4540–90500–1000 cyclesProven cycle, driven by CSPCorrosive chemicals, expensive materials[16, 17]
HyS Cycle40–5060–110500–1000 cyclesCombines electrochemical and thermochemicalRequires both heat and electricity[18, 19]
Ca-Br Cycle25–3530–70500–1000 cyclesLower temperature processLower efficiency, complex handling[20]
Catalytic Coatings on Metal Oxides30–5050–1001000 cyclesImproves reaction rates, reduces temperaturesHigh cost of precious metals[29, 30]
Optimizing Particle Size30–5050–1001000 cyclesIncreases surface area, enhances reactionsScaling up for industrial use[31]
Reactor Configurations (Fluidized Bed, Rotating Particle)35–5560–1201500 cyclesImproves heat and mass transferComplex engineering for scaling[32]
High-Performance Materials (Ceramics and Refractory Metals)30–5050–1001000 cyclesHigh thermal resistance, Withstands extreme temperaturesExpensive, prone to failure under cycling, oxidation concerns[33, 34]
Coating Technologies to Protect Metal Oxides30–5050–1001200 cyclesPrevents degradation and sinteringRequires robust materials for high temps[35]
Parabolic Troughs35–5060–1101500 cyclesEfficient solar concentrationRequires large solar field[36]
Central Tower Receivers40–6070–1302000 cyclesMaximizes solar concentrationRequires consistent sunlight[37]
Recuperative Heat Exchangers30–5050–1001200 cyclesImproves energy efficiencyDesign complexity[38]
Membrane-Based Separation Systems (Ceramic Membranes, Perovskite Oxides)30–5050–1001500 cyclesEfficient hydrogen separationHigh manufacturing cost[39–41]
Rotating Particle Reactors and Fluidized Bed35–5560–1201500 cyclesOptimizes reaction conditionsComplex integration and design[42, 43]
Nanostructured Metal Oxides and Composite Materials35–5560–1201500 cyclesEnhances catalytic propertiesExpensive to scale up[22, 23, 27]

2.1.1 Mechanisms of cyclic redox reactions

At the core of thermochemical hydrogen production are cyclic redox reactions involving metal oxides, which alternately undergo reduction at high temperatures and re-oxidation by water to produce hydrogen. A widely studied example is the ZnO/Zn cycle (as shown in Fig. 3), which operates through the two-step reactions [21]:

A schematic diagram illustrating the two-step thermochemical cycle for hydrogen production using zinc oxide (ZnO). Reduction step: Solar energy is used at 2000°C to decompose ZnO into zinc (Zn) and oxygen (O₂). Oxidation step: At 400°C, Zn reacts with water (H₂O) to regenerate ZnO and produce hydrogen gas (H₂).
Figure 3.

A depiction of the two-step ZnO/Zn thermochemical cycle.

Thermal reduction: ZnO → Zn + 1/2O2 (at T > 2000°C), during this step, solar energy is concentrated to achieve the high temperatures (>2000°C) required to reduce zinc oxide (ZnO) to metallic zinc and oxygen. Water splitting (hydrolysis): Zn + H2O → ZnO + H2 (at ~ 400–600°C), in this stage, the reduced zinc reacts with water vapor at lower temperatures (~400–600°C) to regenerate ZnO and produce hydrogen gas.

This two-step cycle effectively converts solar thermal energy into chemical energy in the form of hydrogen. However, achieving and maintaining the necessary high temperatures (>2000°C) present significant engineering challenges. Advanced solar concentrator technologies and reactor designs are required to focus and retain heat with minimal losses [44]. The thermodynamics of these reactions determine the feasibility of thermochemical cycles. In other words, the energy required to break chemical bonds and the energy released during the reactions determine whether the process can efficiently produce hydrogen under given conditions, such as temperature and pressure. If the energy input needed is too high, the process may be inefficient or impractical for large-scale hydrogen production. For example, the reduction of metal oxides like ZnO requires substantial input energy, as the Gibbs free energy of the reaction is highly endothermic. The efficiency of this process depends on the ability to minimize heat loss during the high-temperature reduction phase and ensure effective heat recovery between cycles [23]. In addition to ZnO, other metal oxide pairs, such as CeO2/Ce2O3 [24, 26] and Fe3O4/FeO [25], have been investigated due to their lower temperature requirements and favorable thermodynamic properties. However, these cycles typically exhibit lower hydrogen production efficiencies compared to the ZnO/Zn cycle, in which the exploration of alternative materials is an ongoing research focus aimed at balancing efficiency, cost, and operational temperatures.

2.1.2 Developing methods of thermochemical hydrogen production

The reaction kinetics of the redox cycle significantly influence the overall efficiency of thermochemical hydrogen production. For example, the kinetic limitations in the ZnO reduction step stem from the slow rate of oxygen release at high temperatures, which can reduce overall cycle efficiency. Researchers are investigating various strategies to enhance reaction rates, such as using catalytic coatings on metal oxides [29, 30], optimizing particle size to increase surface area [31], and developing novel reactor configurations that improve heat and mass transfer [32].

Material degradation at high temperatures is another critical factor. The extreme thermal conditions during the reduction phase can lead to material degradation, which decreases the surface area available for reactions, thus lowering efficiency over repeated cycles [45]. Furthermore, reactor components and materials must be resistant to high-temperature oxidation and thermal stress [46]. High-performance materials, such as ceramics [33] and refractory metals [34], are being explored to withstand the demanding conditions in thermochemical reactors. Coating technologies [35] that can protect metal oxides from sintering or degradation are also being developed.

Another challenge in thermochemical hydrogen production is improving the efficiency of heat transfer within the system. The need for high temperatures and the difficulty of maintaining these temperatures uniformly across the reaction environment often lead to heat losses [37]. Advanced solar concentrators capable of achieving high flux densities, such as parabolic troughs [47] and central tower receivers [36], have been employed to address this issue, but additional strategies are required to optimize heat utilization.

Heat recovery systems can significantly enhance overall cycle efficiency by capturing excess thermal energy from the high-temperature reduction phase and redirecting it to the subsequent water-splitting phase [38]. The design of recuperative heat exchangers and thermal energy storage systems [37] is a critical area of research that seeks to reduce the energy input required for each cycle, thereby improving the net solar-to-hydrogen efficiency.

Additionally, the separation of hydrogen and oxygen during the process is essential to avoid recombination [48], which can reduce the yield of hydrogen and introduce safety concerns. One promising approach is the use of membrane-based separation systems [39] that allow for the selective removal of oxygen during the reduction phase, thereby preventing recombination and ensuring high-purity hydrogen production. Recent advances in ceramic membranes [40] and perovskite oxides [41] have shown potential in this area, but their scalability and durability under high-temperature conditions require further investigation.

2.1.3 Efficiency and system limitations

Several factors limit the current efficiency of thermochemical hydrogen production systems. First, the high operational temperatures necessitate advanced materials and reactor designs, increasing the capital and operational costs. Second, heat losses due to imperfect insulation and inefficient heat transfer mechanisms reduce the overall efficiency of the system. Third, the slow kinetics of oxygen release during the reduction step hinder the rate of hydrogen production, necessitating the exploration of catalysts or optimized material structures. In response to these challenges, novel reactor designs [12] are being developed that focus on maximizing the concentration and retention of solar energy while minimizing thermal losses. For instance, rotating particle reactors [42] have been proposed to enhance the exposure of metal oxide particles to solar radiation, improving the efficiency of the reduction reaction. Similarly, fluidized bed reactors [43] offer the potential to improve heat transfer and reaction kinetics by ensuring continuous mixing and uniform heating of the reactants. Advanced material engineering also plays a critical role in addressing these limitations. The development of nanostructured metal oxides [22, 27] and composite materials [23] that can withstand extreme temperatures without degradation offers a promising pathway to improving the longevity and efficiency of thermochemical systems.

2.2 Photochemical hydrogen production

Photochemical hydrogen production utilizes solar energy to drive water-splitting reactions through photocatalysis, which has garnered significant interest due to its potential to provide a direct, low-cost pathway for converting sunlight into hydrogen [49]. However, the efficiency of current systems is limited by the properties of available photocatalysts, which typically absorb only a small portion of the solar spectrum, and by challenges related to charge separation and transfer within the catalyst material [50]. Advances in photocatalyst development and system optimization are therefore crucial to improving the overall efficiency of photochemical hydrogen production. Table 2 offers a detailed summary of photochemical hydrogen production methods, highlighting their efficiency, production rates, material durability, key advantages, and associated challenges.

Table 2.

Summary of photochemical hydrogen production methods, efficiency, production rate, material durability, and challenges.

MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
TiO25–1010–201000 cyclesStable, abundantOnly absorbs UV light[51]
Transition Metal Doping (Fe, Co, Cr)8–1515–25500–800 cyclesEnhances visible light absorptionDoping levels must be controlled[52–56]
Nonmetal Doping (N, S, C)10–1815–30600–900 cyclesCost-effective improvement for visible lightBalancing doping levels for performance[57–60]
TiO2 Nanostructuring (Nanotubes, Nanosheets)15–2020–401000 cyclesIncreases surface area and improves efficiencyFabrication challenges, scaling issues[61–64]
TiO2 Core-Shell Structures18–2220–50700–1000 cyclesEnhances charge separationComplex fabrication, costly[53, 65]
Transition Metal Oxides (ZnO, WO3, Fe2O3)8–1210–20500–700 cyclesImproved visible light performanceStability under light and pH conditions[66–70]
Oxide Heterojunctions (ZnO with rGO)15–2520–50600–900 cyclesImproved charge separation, better light utilizationComplex synthesis, stability concerns[71–73]
Perovskite Materials20–2550–80<300 cyclesHigh efficiency, tunable propertiesMoisture and heat sensitivity[74]
MOFs5–1210–25<300 cyclesLarge surface area, tunable propertiesExpensive, fragile[75]
Band Gap Engineering (Quantum Confinement, Defect Engineering)15–2520–50600–900 cyclesImproves light absorptionDifficult to control defects and structural integrity[76–78]
Electron-Hole Recombination (Pt, Ni)20–3050–1001000 cyclesReduces electron-hole recombinationHigh cost of Pt, Ni[79–82]
Plasmonic Nanoparticles (Au, Ag)20–3040–80600–800 cyclesEnhanced light absorption via plasmonicsHigh cost of Au, Ag[83–85]
Z-Scheme Photocatalysts25–3580–1501200 cyclesMimics natural photosynthesis, high efficiencyComplex synthesis, integration challenges[86, 87]
Layered Photocatalysts (2D materials—Graphene, MoS2)25–3560–120800–1000 cyclesHigh surface area, excellent charge mobilityScaling up production is challenging[71, 83]
Nanostructures15–2530–501000 cyclesEnhanced reaction ratesFabrication complexity, scaling[60–64, 69, 70, 77, 85]
Suspension Reactors10–2025–451000 cyclesMaintains catalyst in optimal conditionsOptimizing light penetration, energy losses[88]
MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
TiO25–1010–201000 cyclesStable, abundantOnly absorbs UV light[51]
Transition Metal Doping (Fe, Co, Cr)8–1515–25500–800 cyclesEnhances visible light absorptionDoping levels must be controlled[52–56]
Nonmetal Doping (N, S, C)10–1815–30600–900 cyclesCost-effective improvement for visible lightBalancing doping levels for performance[57–60]
TiO2 Nanostructuring (Nanotubes, Nanosheets)15–2020–401000 cyclesIncreases surface area and improves efficiencyFabrication challenges, scaling issues[61–64]
TiO2 Core-Shell Structures18–2220–50700–1000 cyclesEnhances charge separationComplex fabrication, costly[53, 65]
Transition Metal Oxides (ZnO, WO3, Fe2O3)8–1210–20500–700 cyclesImproved visible light performanceStability under light and pH conditions[66–70]
Oxide Heterojunctions (ZnO with rGO)15–2520–50600–900 cyclesImproved charge separation, better light utilizationComplex synthesis, stability concerns[71–73]
Perovskite Materials20–2550–80<300 cyclesHigh efficiency, tunable propertiesMoisture and heat sensitivity[74]
MOFs5–1210–25<300 cyclesLarge surface area, tunable propertiesExpensive, fragile[75]
Band Gap Engineering (Quantum Confinement, Defect Engineering)15–2520–50600–900 cyclesImproves light absorptionDifficult to control defects and structural integrity[76–78]
Electron-Hole Recombination (Pt, Ni)20–3050–1001000 cyclesReduces electron-hole recombinationHigh cost of Pt, Ni[79–82]
Plasmonic Nanoparticles (Au, Ag)20–3040–80600–800 cyclesEnhanced light absorption via plasmonicsHigh cost of Au, Ag[83–85]
Z-Scheme Photocatalysts25–3580–1501200 cyclesMimics natural photosynthesis, high efficiencyComplex synthesis, integration challenges[86, 87]
Layered Photocatalysts (2D materials—Graphene, MoS2)25–3560–120800–1000 cyclesHigh surface area, excellent charge mobilityScaling up production is challenging[71, 83]
Nanostructures15–2530–501000 cyclesEnhanced reaction ratesFabrication complexity, scaling[60–64, 69, 70, 77, 85]
Suspension Reactors10–2025–451000 cyclesMaintains catalyst in optimal conditionsOptimizing light penetration, energy losses[88]
Table 2.

Summary of photochemical hydrogen production methods, efficiency, production rate, material durability, and challenges.

MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
TiO25–1010–201000 cyclesStable, abundantOnly absorbs UV light[51]
Transition Metal Doping (Fe, Co, Cr)8–1515–25500–800 cyclesEnhances visible light absorptionDoping levels must be controlled[52–56]
Nonmetal Doping (N, S, C)10–1815–30600–900 cyclesCost-effective improvement for visible lightBalancing doping levels for performance[57–60]
TiO2 Nanostructuring (Nanotubes, Nanosheets)15–2020–401000 cyclesIncreases surface area and improves efficiencyFabrication challenges, scaling issues[61–64]
TiO2 Core-Shell Structures18–2220–50700–1000 cyclesEnhances charge separationComplex fabrication, costly[53, 65]
Transition Metal Oxides (ZnO, WO3, Fe2O3)8–1210–20500–700 cyclesImproved visible light performanceStability under light and pH conditions[66–70]
Oxide Heterojunctions (ZnO with rGO)15–2520–50600–900 cyclesImproved charge separation, better light utilizationComplex synthesis, stability concerns[71–73]
Perovskite Materials20–2550–80<300 cyclesHigh efficiency, tunable propertiesMoisture and heat sensitivity[74]
MOFs5–1210–25<300 cyclesLarge surface area, tunable propertiesExpensive, fragile[75]
Band Gap Engineering (Quantum Confinement, Defect Engineering)15–2520–50600–900 cyclesImproves light absorptionDifficult to control defects and structural integrity[76–78]
Electron-Hole Recombination (Pt, Ni)20–3050–1001000 cyclesReduces electron-hole recombinationHigh cost of Pt, Ni[79–82]
Plasmonic Nanoparticles (Au, Ag)20–3040–80600–800 cyclesEnhanced light absorption via plasmonicsHigh cost of Au, Ag[83–85]
Z-Scheme Photocatalysts25–3580–1501200 cyclesMimics natural photosynthesis, high efficiencyComplex synthesis, integration challenges[86, 87]
Layered Photocatalysts (2D materials—Graphene, MoS2)25–3560–120800–1000 cyclesHigh surface area, excellent charge mobilityScaling up production is challenging[71, 83]
Nanostructures15–2530–501000 cyclesEnhanced reaction ratesFabrication complexity, scaling[60–64, 69, 70, 77, 85]
Suspension Reactors10–2025–451000 cyclesMaintains catalyst in optimal conditionsOptimizing light penetration, energy losses[88]
MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
TiO25–1010–201000 cyclesStable, abundantOnly absorbs UV light[51]
Transition Metal Doping (Fe, Co, Cr)8–1515–25500–800 cyclesEnhances visible light absorptionDoping levels must be controlled[52–56]
Nonmetal Doping (N, S, C)10–1815–30600–900 cyclesCost-effective improvement for visible lightBalancing doping levels for performance[57–60]
TiO2 Nanostructuring (Nanotubes, Nanosheets)15–2020–401000 cyclesIncreases surface area and improves efficiencyFabrication challenges, scaling issues[61–64]
TiO2 Core-Shell Structures18–2220–50700–1000 cyclesEnhances charge separationComplex fabrication, costly[53, 65]
Transition Metal Oxides (ZnO, WO3, Fe2O3)8–1210–20500–700 cyclesImproved visible light performanceStability under light and pH conditions[66–70]
Oxide Heterojunctions (ZnO with rGO)15–2520–50600–900 cyclesImproved charge separation, better light utilizationComplex synthesis, stability concerns[71–73]
Perovskite Materials20–2550–80<300 cyclesHigh efficiency, tunable propertiesMoisture and heat sensitivity[74]
MOFs5–1210–25<300 cyclesLarge surface area, tunable propertiesExpensive, fragile[75]
Band Gap Engineering (Quantum Confinement, Defect Engineering)15–2520–50600–900 cyclesImproves light absorptionDifficult to control defects and structural integrity[76–78]
Electron-Hole Recombination (Pt, Ni)20–3050–1001000 cyclesReduces electron-hole recombinationHigh cost of Pt, Ni[79–82]
Plasmonic Nanoparticles (Au, Ag)20–3040–80600–800 cyclesEnhanced light absorption via plasmonicsHigh cost of Au, Ag[83–85]
Z-Scheme Photocatalysts25–3580–1501200 cyclesMimics natural photosynthesis, high efficiencyComplex synthesis, integration challenges[86, 87]
Layered Photocatalysts (2D materials—Graphene, MoS2)25–3560–120800–1000 cyclesHigh surface area, excellent charge mobilityScaling up production is challenging[71, 83]
Nanostructures15–2530–501000 cyclesEnhanced reaction ratesFabrication complexity, scaling[60–64, 69, 70, 77, 85]
Suspension Reactors10–2025–451000 cyclesMaintains catalyst in optimal conditionsOptimizing light penetration, energy losses[88]

2.2.1 Photocatalyst classes

Photocatalysts play a central role in photochemical hydrogen production, driving the water-splitting reaction by absorbing light, generating charge carriers (electrons and holes), and facilitating redox reactions. These catalysts can be broadly categorized into three classes:

  1. Single-component photocatalysts

Single-component materials, such as titanium dioxide (TiO2), ZnO, tungsten trioxide (WO3), and hematite (Fe2O3), are widely studied due to their simplicity and well-understood properties. TiO2, for example, is highly stable, cost-effective, and exhibits significant photocatalytic activity under ultraviolet (UV) light. However, its wide bandgap (~3.2 eV) restricts absorption to the UV spectrum, which constitutes only ~4% of solar radiation [51].

To extend absorption into the visible range, researchers have employed doping strategies, introducing elements such as Fe [52–54], Co [55], Cr [56] (transition metals) or N [57], S [58, 59], and C [60] (nonmetals). Although doping can reduce the bandgap, it often introduces recombination centers that decrease overall efficiency [51]. Nanostructuring TiO2 into 1D (nanotubes [61, 62]) or 2D (nanosheets [63, 64]) forms improves charge separation and transport by providing shorter diffusion paths for charge carriers. Core-shell structures, where TiO2 is coated with other materials, offer additional advantages by protecting the catalyst from recombination and maintaining high stability [53, 65].

Other single-component materials, such as ZnO [66, 67], WO3 [68], and Fe2O3 [69, 70], feature narrower bandgaps for visible light absorption but suffer from poor charge transport properties and low quantum efficiencies. To address these limitations, heterojunction designs combining these materials with others have been explored [89].

  1. Type-II heterojunction photocatalysts

Type-II heterojunctions involve interfaces between two materials with complementary band structures. These structures promote efficient charge separation by confining electrons and holes to different regions. For example, coupling ZnO with reduced graphene oxide (rGO) improves electron mobility and extends light absorption into the visible range [71–73]. These heterostructures show significant promise in improving photocatalytic performance by leveraging synergies between their components.

  1. Z-Scheme photocatalysts

Inspired by natural photosynthesis, Z-scheme systems use two photocatalysts with complementary bandgaps [87], one driving the oxygen evolution reaction (OER) and the other driving the hydrogen evolution reaction (HER). An electron mediator transfers charge between the two photocatalysts, allowing for better utilization of the solar spectrum. Examples include systems incorporating plasmonic nanoparticles or 2D materials like graphene [90] and molybdenum disulfide (MoS2) [71, 83]. Z-scheme designs are particularly promising for overcoming the limitations of single-component and type-II heterojunction systems, achieving higher overall efficiency by separating reaction sites [86].

2.2.2 Bandgap engineering and light absorption mechanisms

A fundamental aspect of photochemical hydrogen production is the bandgap of the photocatalyst (as shown in Fig. 4), which defines its ability to absorb light and generate the charge carriers necessary for water splitting. The bandgap is the energy difference between the valence band (occupied by electrons) and the conduction band (unoccupied energy levels). For a photocatalyst to efficiently drive water splitting, its bandgap must allow absorption of a significant portion of the solar spectrum and generate electrons and holes with sufficient energy for the redox reactions. The minimum theoretical bandgap required for water splitting is ~1.23 eV, corresponding to the Gibbs free energy of the reaction [49]. However, in practical systems, photocatalysts typically have bandgaps in the range of 1.5–3 eV to account for additional energy losses due to recombination, overpotential, and energy dissipation at the electrode surface [76].

A schematic diagram of photocatalytic water splitting using a semiconductor. Solar energy excites electrons in the semiconductor, generating electron-hole pairs. Electron pathway: Electrons reduce protons (H⁺) to produce hydrogen gas (H₂). Hole pathway: Holes oxidize water (H₂O) to generate oxygen gas (O₂). Recombination: Some electrons and holes recombine, reducing efficiency.
Figure 4.

Diagram illustrating the fundamental mechanism of the photocatalytic water-splitting system.

The light absorption mechanism begins when photons with energy equal to or greater than the bandgap excite electrons from the valence band to the conduction band, leaving behind holes in the valence band. This process generates electron-hole pairs, or excitons, which are the primary drivers of the HER and OER. Photocatalysts with smaller bandgaps (<1.5 eV) can absorb a broader range of the solar spectrum, including visible and near-infrared light, but may lack the energy required to efficiently drive water splitting. Conversely, larger bandgaps (>3 eV) provide sufficient energy for redox reactions but are limited to absorbing UV light, which constitutes only a small fraction (~4%) of the solar spectrum [49].

Bandgap engineering techniques are essential for optimizing light absorption in photocatalysts, enabling improved performance in photochemical hydrogen production. One approach is quantum confinement [77], which involves reducing the dimensions of materials to the nanoscale, thereby altering their electronic properties, increasing the bandgap, and enhancing charge dynamics. Another strategy is defect engineering [78], where controlled defects are introduced into the material’s structure, creating localized energy states that improve light absorption and charge carrier mobility. Additionally, the incorporation of plasmonic nanoparticles, such as gold (Au) [83] or silver (Ag) [84, 85], enhances light absorption by leveraging localized surface plasmon resonance, which amplifies electromagnetic fields near the catalyst surface and boosts the generation of electron-hole pairs. These techniques collectively contribute to more efficient utilization of the solar spectrum for water splitting.

2.2.3 Optimization of photocatalytic systems

Improving the solar-to-hydrogen conversion efficiency requires not only advances in photocatalyst materials but also system-level optimization. Suspension reactors [88], where photocatalysts are dispersed in a liquid medium, enhance the interaction between light, water, and catalysts, increasing reaction rates. Layered 2D materials, such as graphene [90] and MoS2 [71, 83], provide broad light absorption and efficient charge transport pathways. Additionally, high surface area nanostructured materials [59–64, 69, 70, 77, 85] offer more active sites for water splitting, further boosting efficiency.

Visible light utilization remains a key focus, as UV light accounts for only a small fraction of solar energy. Researchers [91, 92] are developing photocatalysts with broad-spectrum absorption, multifunctional designs, and robust frameworks to improve light harvesting and stability under real-world conditions. Metal-organic frameworks (MOFs) [93] and perovskites [74], while not photocatalysts themselves, serve as scaffolds for active components, offering tunable properties and controlled environments for hydrogen evolution [75]. However, their stability in aqueous environments [94], potential toxicity due to lead content [95], and sensitivity to moisture [96] remain major challenges. Addressing these challenges requires innovative solutions, such as developing lead-free perovskites [97], enhancing stability through surface passivation techniques [98], encapsulating the materials to protect them from environmental exposure [99], and incorporating hybrid organic-inorganic structures to improve resistance to moisture and aqueous conditions [100].

2.3 Biological hydrogen production

Biological hydrogen production is a promising approach that utilizes the natural metabolic processes of microorganisms such as algae, cyanobacteria, and photosynthetic bacteria to produce hydrogen from sunlight and water [101]. While this method has the potential to offer a sustainable and low-cost solution for hydrogen production, its efficiency and scalability remain limited by several biochemical and engineering challenges [102]. Advances in understanding the metabolic pathways involved in hydrogen production, as well as the development of efficient bioreactor systems, are critical to improving the scalability and feasibility of this technology. Table 3 presents a comprehensive overview of biological hydrogen production methods, emphasizing their efficiency, production rates, material durability, key benefits, and challenges.

Table 3.

Summary of biological hydrogen production methods, efficiency, production rate, material durability, and challenges.

MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
Nitrogenase5–100.5–2<100 cyclesCapable of producing hydrogen under nitrogen-limited conditionsRequires large amounts of ATP, sensitive to oxygen[103]
Hydrogenase15–255–10200–400 cyclesMore efficient under anoxic conditionsSensitive to oxygen, limiting efficiency[104]
Genetic Modification (Algae and Cyanobacteria)30–3010–15200–400 cyclesEnhances hydrogen production in photosynthetic organismsMaintaining genetic stability, complexity of engineering[105]
Electron Channeling to Hydrogenase or Nitrogenase30–4015–25300 cyclesImproves efficiency of hydrogen productionBalancing electron flow and preventing competition[106]
Knocking Out Competing Pathways (Calvin cycle and respiration)25–3510–20300–500 cyclesIncreases electron flow to hydrogen productionCan reduce organism fitness and growth rates[107, 108]
Synthetic Biology (Nanomaterials or Artificial Electron Donors)40–5020–30200–400 cyclesEnhances electron transfer to enzymesComplex integration of nanomaterials with biology[109, 110]
Hybrid Systems Incorporating Metal Nanoparticles45–5525–35300–500 cyclesIncreases catalytic efficiencyHigh cost, scaling challenges[111]
Site-Directed Mutagenesis40–5020–30300–400 cyclesImproves enzyme functionalityUnintended effects on enzyme stability[112]
Designing Enzyme Electron Transfer Pathways50–6025–40500 cyclesBoosts electron transfer efficiencyMaintaining stability and functionality in modified pathways[113]
Cellular Strategies to Limit Hydrogenase Exposure to Oxygen30–4015–25200 cyclesReduces oxygen inhibition of hydrogenaseDifficult to maintain in fluctuating environments[114]
Optical Fiber-Based Reactors and Internal Light Sources40–5020–35300–500 cyclesIncreases light availability for photosynthetic hydrogen productionCostly, difficult to scale[115, 116]
Removing Oxygen and Supplying CO230–4015–25300–400 cyclesMaintains favorable conditions for hydrogen productionComplex and costly membrane systems[117]
Mixed Microbial Cultures35–4520–30300–500 cyclesCombines complementary pathways for higher efficiencyMaintaining microbial community stability[118]
Optimizing Light-to-Hydrogen Conversion Efficiency45–5525–40500 cyclesMaximizes hydrogen yieldDifficult to achieve under real-world conditions[119]
Oxygen Management (Membrane-Based Gas Separators)40–5020–35500 cyclesPrevents oxygen inhibitionDeveloping cost-effective separation technologies[39]
MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
Nitrogenase5–100.5–2<100 cyclesCapable of producing hydrogen under nitrogen-limited conditionsRequires large amounts of ATP, sensitive to oxygen[103]
Hydrogenase15–255–10200–400 cyclesMore efficient under anoxic conditionsSensitive to oxygen, limiting efficiency[104]
Genetic Modification (Algae and Cyanobacteria)30–3010–15200–400 cyclesEnhances hydrogen production in photosynthetic organismsMaintaining genetic stability, complexity of engineering[105]
Electron Channeling to Hydrogenase or Nitrogenase30–4015–25300 cyclesImproves efficiency of hydrogen productionBalancing electron flow and preventing competition[106]
Knocking Out Competing Pathways (Calvin cycle and respiration)25–3510–20300–500 cyclesIncreases electron flow to hydrogen productionCan reduce organism fitness and growth rates[107, 108]
Synthetic Biology (Nanomaterials or Artificial Electron Donors)40–5020–30200–400 cyclesEnhances electron transfer to enzymesComplex integration of nanomaterials with biology[109, 110]
Hybrid Systems Incorporating Metal Nanoparticles45–5525–35300–500 cyclesIncreases catalytic efficiencyHigh cost, scaling challenges[111]
Site-Directed Mutagenesis40–5020–30300–400 cyclesImproves enzyme functionalityUnintended effects on enzyme stability[112]
Designing Enzyme Electron Transfer Pathways50–6025–40500 cyclesBoosts electron transfer efficiencyMaintaining stability and functionality in modified pathways[113]
Cellular Strategies to Limit Hydrogenase Exposure to Oxygen30–4015–25200 cyclesReduces oxygen inhibition of hydrogenaseDifficult to maintain in fluctuating environments[114]
Optical Fiber-Based Reactors and Internal Light Sources40–5020–35300–500 cyclesIncreases light availability for photosynthetic hydrogen productionCostly, difficult to scale[115, 116]
Removing Oxygen and Supplying CO230–4015–25300–400 cyclesMaintains favorable conditions for hydrogen productionComplex and costly membrane systems[117]
Mixed Microbial Cultures35–4520–30300–500 cyclesCombines complementary pathways for higher efficiencyMaintaining microbial community stability[118]
Optimizing Light-to-Hydrogen Conversion Efficiency45–5525–40500 cyclesMaximizes hydrogen yieldDifficult to achieve under real-world conditions[119]
Oxygen Management (Membrane-Based Gas Separators)40–5020–35500 cyclesPrevents oxygen inhibitionDeveloping cost-effective separation technologies[39]
Table 3.

Summary of biological hydrogen production methods, efficiency, production rate, material durability, and challenges.

MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
Nitrogenase5–100.5–2<100 cyclesCapable of producing hydrogen under nitrogen-limited conditionsRequires large amounts of ATP, sensitive to oxygen[103]
Hydrogenase15–255–10200–400 cyclesMore efficient under anoxic conditionsSensitive to oxygen, limiting efficiency[104]
Genetic Modification (Algae and Cyanobacteria)30–3010–15200–400 cyclesEnhances hydrogen production in photosynthetic organismsMaintaining genetic stability, complexity of engineering[105]
Electron Channeling to Hydrogenase or Nitrogenase30–4015–25300 cyclesImproves efficiency of hydrogen productionBalancing electron flow and preventing competition[106]
Knocking Out Competing Pathways (Calvin cycle and respiration)25–3510–20300–500 cyclesIncreases electron flow to hydrogen productionCan reduce organism fitness and growth rates[107, 108]
Synthetic Biology (Nanomaterials or Artificial Electron Donors)40–5020–30200–400 cyclesEnhances electron transfer to enzymesComplex integration of nanomaterials with biology[109, 110]
Hybrid Systems Incorporating Metal Nanoparticles45–5525–35300–500 cyclesIncreases catalytic efficiencyHigh cost, scaling challenges[111]
Site-Directed Mutagenesis40–5020–30300–400 cyclesImproves enzyme functionalityUnintended effects on enzyme stability[112]
Designing Enzyme Electron Transfer Pathways50–6025–40500 cyclesBoosts electron transfer efficiencyMaintaining stability and functionality in modified pathways[113]
Cellular Strategies to Limit Hydrogenase Exposure to Oxygen30–4015–25200 cyclesReduces oxygen inhibition of hydrogenaseDifficult to maintain in fluctuating environments[114]
Optical Fiber-Based Reactors and Internal Light Sources40–5020–35300–500 cyclesIncreases light availability for photosynthetic hydrogen productionCostly, difficult to scale[115, 116]
Removing Oxygen and Supplying CO230–4015–25300–400 cyclesMaintains favorable conditions for hydrogen productionComplex and costly membrane systems[117]
Mixed Microbial Cultures35–4520–30300–500 cyclesCombines complementary pathways for higher efficiencyMaintaining microbial community stability[118]
Optimizing Light-to-Hydrogen Conversion Efficiency45–5525–40500 cyclesMaximizes hydrogen yieldDifficult to achieve under real-world conditions[119]
Oxygen Management (Membrane-Based Gas Separators)40–5020–35500 cyclesPrevents oxygen inhibitionDeveloping cost-effective separation technologies[39]
MethodsEfficiency (%)Production Rate (mg H2/l/h)Material DurabilityAdvantagesChallengesCitations
Nitrogenase5–100.5–2<100 cyclesCapable of producing hydrogen under nitrogen-limited conditionsRequires large amounts of ATP, sensitive to oxygen[103]
Hydrogenase15–255–10200–400 cyclesMore efficient under anoxic conditionsSensitive to oxygen, limiting efficiency[104]
Genetic Modification (Algae and Cyanobacteria)30–3010–15200–400 cyclesEnhances hydrogen production in photosynthetic organismsMaintaining genetic stability, complexity of engineering[105]
Electron Channeling to Hydrogenase or Nitrogenase30–4015–25300 cyclesImproves efficiency of hydrogen productionBalancing electron flow and preventing competition[106]
Knocking Out Competing Pathways (Calvin cycle and respiration)25–3510–20300–500 cyclesIncreases electron flow to hydrogen productionCan reduce organism fitness and growth rates[107, 108]
Synthetic Biology (Nanomaterials or Artificial Electron Donors)40–5020–30200–400 cyclesEnhances electron transfer to enzymesComplex integration of nanomaterials with biology[109, 110]
Hybrid Systems Incorporating Metal Nanoparticles45–5525–35300–500 cyclesIncreases catalytic efficiencyHigh cost, scaling challenges[111]
Site-Directed Mutagenesis40–5020–30300–400 cyclesImproves enzyme functionalityUnintended effects on enzyme stability[112]
Designing Enzyme Electron Transfer Pathways50–6025–40500 cyclesBoosts electron transfer efficiencyMaintaining stability and functionality in modified pathways[113]
Cellular Strategies to Limit Hydrogenase Exposure to Oxygen30–4015–25200 cyclesReduces oxygen inhibition of hydrogenaseDifficult to maintain in fluctuating environments[114]
Optical Fiber-Based Reactors and Internal Light Sources40–5020–35300–500 cyclesIncreases light availability for photosynthetic hydrogen productionCostly, difficult to scale[115, 116]
Removing Oxygen and Supplying CO230–4015–25300–400 cyclesMaintains favorable conditions for hydrogen productionComplex and costly membrane systems[117]
Mixed Microbial Cultures35–4520–30300–500 cyclesCombines complementary pathways for higher efficiencyMaintaining microbial community stability[118]
Optimizing Light-to-Hydrogen Conversion Efficiency45–5525–40500 cyclesMaximizes hydrogen yieldDifficult to achieve under real-world conditions[119]
Oxygen Management (Membrane-Based Gas Separators)40–5020–35500 cyclesPrevents oxygen inhibitionDeveloping cost-effective separation technologies[39]

2.3.1 Mechanisms of enzymes hydrogen production

Biological hydrogen production relies on specific enzymes within microorganisms that catalyze the conversion of water or organic substrates into hydrogen, in which hydrogenase and nitrogenase play crucial roles in the hydrogen evolution process.

Hydrogenase enzymes are widely distributed among microorganisms and catalyze the reversible reduction of protons to molecular hydrogen [101]. These enzymes exist in three main types [104] based on their active metal centers: [FeFe]-hydrogenases, [NiFe]-hydrogenases, and [Fe]-hydrogenases; their structures and active sites are shown in Fig. 5. The [FeFe]-hydrogenases and [NiFe]-hydrogenases are the most extensively studied, as they perform similar functions but differ in their active sites. [FeFe]-hydrogenases are typically associated with higher catalytic activity, while [NiFe]-hydrogenases are more common and often more robust under varying conditions [120]. Both types utilize electrons generated during photosynthesis or the metabolic degradation of organic compounds to reduce protons into hydrogen gas. However, a significant limitation of these enzymes is their sensitivity to oxygen, which can irreversibly inactivate the enzyme by damaging its metal centers. This sensitivity poses a major bottleneck in biological hydrogen production, particularly in photosynthetic systems where oxygen is a byproduct [121]. The third type, [Fe]-hydrogenases, contains a single iron atom at the active site without iron-sulfur clusters, distinguishing it from the other two types. These enzymes operate via a fundamentally different enzymatic mechanism, primarily catalyzing hydrogen production in methanogenic archaea [122]. Although less studied, [Fe]-hydrogenases contribute to the diversity of enzymatic pathways for hydrogen evolution and may offer unique advantages in specific microbial systems.

A diagram illustrating the structures of three types of hydrogenase enzymes:(a) [NiFe]-hydrogenase: Shows a Ni-Fe center with cysteine ligands and additional CO and CN ligands on the Fe atom. (b) [FeFe]-hydrogenase: Features a di-iron cluster with CO and CN ligands, linked to a [4Fe-4S] cluster via a cysteine ligand. (c) [Fe]-hydrogenase: Depicts a mononuclear Fe center coordinated with a sulfur, CO, and nitrogen-based ligands.
Figure 5.

Structure and active site of (a) [NiFe]-hydrogenase, (b) [FeFe]-hydrogenase, and (c) [Fe]-hydrogenase.

Nitrogenase enzymes, found in nitrogen-fixing bacteria and cyanobacteria, also play a role in hydrogen production. Nitrogenase catalyzes the reduction of atmospheric nitrogen (N2) into ammonia (NH3), but hydrogen is produced as a byproduct of this reaction [103]. Unlike hydrogenases, nitrogenase can function in both aerobic and anaerobic conditions, although its hydrogen production efficiency is typically lower than that of hydrogenase. The adenosine triphosphate requirements of nitrogenase, coupled with its dual role in nitrogen fixation and hydrogen evolution, make it less efficient for large-scale hydrogen production compared to dedicated hydrogenases [104].

2.3.2 Metabolic engineering approaches

To overcome the limitations of natural microbial pathways, metabolic engineering and genetic modification approaches have been extensively explored to improve hydrogen yields. One strategy involves modifying the expression levels of hydrogenase or nitrogenase enzymes to enhance hydrogen production [123]. For example, researchers have genetically engineered algae [105] and cyanobacteria [106] to overexpress oxygen-tolerant hydrogenases, thereby improving their hydrogen production capacity under aerobic conditions. Another approach focuses on redirecting the flow of electrons within the microbial cell to favor hydrogen production. In photosynthetic microorganisms, light energy absorbed by photosystems generates high-energy electrons that are transferred to various cellular processes. By manipulating the electron transport chain, it is possible to channel more electrons towards hydrogenase or nitrogenase enzymes, increasing hydrogen output [124], which can be achieved by knocking out competing pathways, such as those involved in carbon fixation (e.g. Calvin cycle [107]) or respiration [108], which otherwise divert electrons away from hydrogen production. In addition, synthetic biology has been used to create hybrid systems that combine microbial cells with synthetic components, such as nanomaterials [109] or artificial electron donors [110], to improve the efficiency of electron transfer and hydrogen evolution. For instance, hybrid systems incorporating metal nanoparticles [111] have been developed to enhance the solar energy capture capabilities of algae and cyanobacteria, thereby increasing the overall efficiency of photosynthetic hydrogen production.

2.3.3 Hydrogenase stability and scalability

A central challenge in biological hydrogen production is improving the stability and efficiency of hydrogenase enzymes, particularly under oxygenic conditions [125]. Recent research has focused on modifying the active site of hydrogenase enzymes to increase their resistance to oxygen [126]. Another major hurdle in enzyme-based hydrogen production is scalability. While hydrogenase and nitrogenase are effective at the laboratory scale, their sensitivity to environmental conditions, particularly oxygen, poses significant challenges for industrial deployment [127]. Site-directed mutagenesis [112] has been employed to alter the metal center of the enzyme structure, reducing its affinity for oxygen while maintaining catalytic activity. Another promising approach involves engineering the electron transfer pathway of the enzyme [113] to increase the rate of hydrogen production, reducing the probability of oxygen-induced inactivation. In addition to enzyme modifications, researchers [114] are exploring cellular strategies that limit the exposure of hydrogenase to oxygen. For example, some cyanobacteria have been engineered to temporally separate photosynthesis and hydrogen production by performing these processes at different times of day [128] (e.g. photosynthesis during the day and hydrogen production at night, when oxygen levels are lower). These efforts are crucial for overcoming the oxygen sensitivity barrier that currently limits the efficiency of biological hydrogen production.

2.3.4 Bioreactor design and scaling challenges

Beyond enzymatic and metabolic improvements, the design of bioreactor systems plays a critical role in determining the overall hydrogen production rate. Bioreactors are engineered environments that support the growth and metabolic activity of microorganisms under controlled conditions [129]. Optimizing these systems for large-scale hydrogen production presents several challenges, particularly in terms of light distribution, nutrient supply, and gas exchange. Light intensity and distribution are key factors in maximizing the efficiency of photosynthetic hydrogen production. In algae and cyanobacteria-based systems, light must be evenly distributed throughout the reactor to ensure that all cells receive adequate illumination for photosynthesis [128]. Traditional flat-panel or tubular photobioreactors often face the issue of light attenuation, where cells near the reactor surface absorb most of the available light, leaving the cells deeper within the reactor with insufficient photon exposure [130]. To address this issue, optical fiber-based reactors [115, 116] and internal light sources [131] have been developed to distribute light more uniformly throughout the culture medium. Additionally, the management of CO2 and oxygen levels is critical for maintaining high hydrogen production rates. In algal systems, CO2 is a key substrate for photosynthesis, and its concentration must be carefully controlled to ensure efficient carbon fixation and electron generation. At the same time, oxygen produced during photosynthesis must be removed from the system to prevent the inhibition of hydrogenase activity [117]. Gas exchange systems that continuously remove oxygen and supply CO2 are therefore essential components of any large-scale bioreactor designed for hydrogen production [130]. Recent advances in biohydrogen production have also explored the use of mixed microbial cultures that combine different species of algae, cyanobacteria, or bacteria to improve hydrogen yields. By employing species with complementary metabolic functions, such as one organism producing hydrogen and another consuming oxygen, these systems can achieve more stable and efficient hydrogen production [118]. However, maintaining the balance between different microbial species in a mixed-culture reactor is challenging, particularly as conditions within the reactor change over time.

2.3.5 Improvement of algae and cyanobacteria biohydrogen production

In controlled laboratory environments, significant progress has been made in improving the efficiency of algal and cyanobacterial hydrogen production. For instance, researchers [119] have optimized the light-to-hydrogen conversion efficiency by adjusting light intensity and wavelength to match the specific absorption characteristics of the photosynthetic machinery. Using light-emitting diode-based illumination systems, which can be tuned to specific wavelengths, scientists have been able to maximize the efficiency of photosystems involved in electron transport and hydrogen evolution. Another important factor is oxygen management, as discussed earlier. In closed photobioreactor systems, advanced membrane-based gas separators [39] are being developed to continuously remove oxygen while maintaining optimal CO2 concentrations. These systems help maintain anaerobic or microaerobic conditions that are favorable for hydrogenase activity, thereby enhancing overall hydrogen production rates. In summary, while biological hydrogen production remains a promising technology, significant challenges must be overcome to scale up these systems for industrial use. Advances in metabolic engineering, enzyme stability, and bioreactor design will be essential for improving the efficiency and economic viability of biohydrogen production. Continued research into the optimization of light harvesting, gas management, and mixed microbial systems will likely play a crucial role in making biological hydrogen production a competitive alternative in the renewable energy landscape.

3. Solar hydrogen storage technologies

One of the key challenges in creating a sustainable hydrogen economy is the efficient and safe storage of hydrogen. The intermittent nature of solar energy necessitates reliable storage technologies to ensure that hydrogen produced via solar methods can be used when needed [132]. Hydrogen can be stored in various forms including compressed gas [133], liquefied hydrogen [134], or chemically bound to materials [135]. This section explores the physical and chemical principles governing hydrogen storage systems, including the thermodynamic and kinetic factors that influence their performance and suitability for various applications. Figures 6 and 7 present a depiction of various hydrogen storage technologies and illustrate hydrogen density and accessibility across different storage systems. Table 4 compares hydrogen storage technologies, highlighting their capacity, energy density, consumption, safety, cost, and challenges.

Table 4.

Comparison of hydrogen storage technologies: capacity, energy density, consumption, safety, cost, and challenges

TechnologyStorage Capacity (kg H2/m³)Energy Density (kWh/kg)Energy Consumption (kWh/kg)Safety (Pressure/Temperature)Cost (USD/kg H2)AdvantagesChallengesCitations
Compressed Gas~20–40~3310–15350–700 bar5–10Well-established, mature technologyLow volumetric density, requires high pressure (350–700 bar)[133]
Liquefied Hydrogen~71~4012–20 (liquefaction)−253°C storage15–20High storage density, ideal for large volumesHigh energy consumption for liquefaction, boil-off losses[134]
Metal Hydrides~100–150~7–155–10 (thermal cycling)~10 bar, 300°C10–20High volumetric storage, reversible storageHeavy, slow release kinetics, thermal management required[136]
MOFs~30–60~335–8 (adsorption/desorption)Low-pressure operation>20Tunable storage properties, lightweightExpensive materials, low energy density[137]
Chemical Storage (complex hydrides)~85–150~5–106–12 (dehydrogenation)~1–10 bar, 100–200°C10–15High hydrogen content, easier to transportComplex handling, energy required for hydrogen release[138]
TechnologyStorage Capacity (kg H2/m³)Energy Density (kWh/kg)Energy Consumption (kWh/kg)Safety (Pressure/Temperature)Cost (USD/kg H2)AdvantagesChallengesCitations
Compressed Gas~20–40~3310–15350–700 bar5–10Well-established, mature technologyLow volumetric density, requires high pressure (350–700 bar)[133]
Liquefied Hydrogen~71~4012–20 (liquefaction)−253°C storage15–20High storage density, ideal for large volumesHigh energy consumption for liquefaction, boil-off losses[134]
Metal Hydrides~100–150~7–155–10 (thermal cycling)~10 bar, 300°C10–20High volumetric storage, reversible storageHeavy, slow release kinetics, thermal management required[136]
MOFs~30–60~335–8 (adsorption/desorption)Low-pressure operation>20Tunable storage properties, lightweightExpensive materials, low energy density[137]
Chemical Storage (complex hydrides)~85–150~5–106–12 (dehydrogenation)~1–10 bar, 100–200°C10–15High hydrogen content, easier to transportComplex handling, energy required for hydrogen release[138]
Table 4.

Comparison of hydrogen storage technologies: capacity, energy density, consumption, safety, cost, and challenges

TechnologyStorage Capacity (kg H2/m³)Energy Density (kWh/kg)Energy Consumption (kWh/kg)Safety (Pressure/Temperature)Cost (USD/kg H2)AdvantagesChallengesCitations
Compressed Gas~20–40~3310–15350–700 bar5–10Well-established, mature technologyLow volumetric density, requires high pressure (350–700 bar)[133]
Liquefied Hydrogen~71~4012–20 (liquefaction)−253°C storage15–20High storage density, ideal for large volumesHigh energy consumption for liquefaction, boil-off losses[134]
Metal Hydrides~100–150~7–155–10 (thermal cycling)~10 bar, 300°C10–20High volumetric storage, reversible storageHeavy, slow release kinetics, thermal management required[136]
MOFs~30–60~335–8 (adsorption/desorption)Low-pressure operation>20Tunable storage properties, lightweightExpensive materials, low energy density[137]
Chemical Storage (complex hydrides)~85–150~5–106–12 (dehydrogenation)~1–10 bar, 100–200°C10–15High hydrogen content, easier to transportComplex handling, energy required for hydrogen release[138]
TechnologyStorage Capacity (kg H2/m³)Energy Density (kWh/kg)Energy Consumption (kWh/kg)Safety (Pressure/Temperature)Cost (USD/kg H2)AdvantagesChallengesCitations
Compressed Gas~20–40~3310–15350–700 bar5–10Well-established, mature technologyLow volumetric density, requires high pressure (350–700 bar)[133]
Liquefied Hydrogen~71~4012–20 (liquefaction)−253°C storage15–20High storage density, ideal for large volumesHigh energy consumption for liquefaction, boil-off losses[134]
Metal Hydrides~100–150~7–155–10 (thermal cycling)~10 bar, 300°C10–20High volumetric storage, reversible storageHeavy, slow release kinetics, thermal management required[136]
MOFs~30–60~335–8 (adsorption/desorption)Low-pressure operation>20Tunable storage properties, lightweightExpensive materials, low energy density[137]
Chemical Storage (complex hydrides)~85–150~5–106–12 (dehydrogenation)~1–10 bar, 100–200°C10–15High hydrogen content, easier to transportComplex handling, energy required for hydrogen release[138]
A diagram of solar hydrogen storage technologies divided into:Storage Methods: Physical (e.g., compressed gas, liquefied hydrogen, carbon-based materials) and chemical (e.g., MOFs, metal hydrides, LOHCs), highlighting advantages and challenges. Mechanisms: Focus on thermodynamics (adsorption/desorption) and kinetics (rate improvements). Applications: Emphasizes cycling stability, thermal management, scalability, and cost.
Figure 6.

An illustration of various hydrogen storage technologies.

A table comparing hydrogen storage methods based on density and accessibility, arranged horizontally: Compression: High-pressure storage (200–700 bar). Liquefaction: Storage at cryogenic temperatures (-253°C). Physisorption: Materials with large surface areas, such as MOFs (e.g., MOF-5, HKUST-1) and carbon-based materials (e.g., CNTs, graphene). Chemisorption: Metal hydrides (e.g., MgH₂, LaNi₅H₆, TiFeH₂). Chemical Compounds: Complex hydrides (e.g., LiAlH₄, NaBH₄, NH₃BH₃) and ammonia storage compounds. LOHCs: Liquid organic hydrogen carriers (e.g., toluene, dibenzyltoluene).
Figure 7.

Hydrogen density and accessibility across various storage systems.

3.1 Physical and chemical hydrogen storage methods

Hydrogen storage materials are a crucial component of solar hydrogen systems, as they allow for the reversible storage and release of hydrogen at practical temperatures and pressures. The development of efficient storage materials involves balancing several factors: storage capacity, kinetics of hydrogen adsorption and desorption, stability under cycling conditions, cost-effectiveness, and sustainability [139]. Current research focuses on improving these properties in metal hydrides, carbon-based materials, and novel chemical storage systems such as MOFs and complex hydrides [137]. Additionally, alternative storage solutions, including liquid organic hydrogen carriers (LOHCs) and synthetic fuels, are being explored to complement material-based hydrogen storage.

Metal hydrides are one of the most well-researched classes of hydrogen storage materials. They store hydrogen by forming metal-hydrogen bonds, often leading to high hydrogen densities at relatively low pressures [136]. The general reaction for hydrogen storage in metal hydrides can be expressed as: M + H2 → MH2, where M represents a metal or alloy that reacts with hydrogen to form a hydride. Metal hydrides, such as MgH2 [140], LaNi5H6 [141], and TiFeH2 [142–144], offer several advantages, including high volumetric hydrogen density and safe, solid-state storage. However, many hydrides require high temperatures for hydrogen release, which limits their practical applications. For instance, magnesium hydride (MgH2) can store up to 7.6 wt.% hydrogen, but its desorption temperature is around 300°C, necessitating substantial energy input for hydrogen release [140]. Recent advancements in metal hydride research have focused on nanostructuring [145–147] and alloying [148, 149] to reduce desorption temperatures and improve reaction kinetics. For example, nanostructured magnesium hydride [145, 146] has demonstrated faster hydrogen uptake and release due to the increased surface area and reduced diffusion path lengths for hydrogen atoms. Additionally, the incorporation of transition metals like Ti [150] or Ni [151] as catalysts has been shown to lower the activation energy for hydrogen desorption, although repeated cycling at high temperatures poses challenges for material durability and recyclability.

Carbon-based materials, including carbon nanotubes (CNTs) [152], graphene [153, 154], and activated carbons [155], offer another route for hydrogen storage, storing hydrogen primarily through physisorption, where hydrogen molecules weakly bind to the surface of the material. While carbon materials typically have lower hydrogen storage capacities compared to metal hydrides, they operate at lower temperatures and pressures, making them more energy-efficient for hydrogen release. Research into doping carbon structures with metals [156, 157] or creating porous carbon composites [158–160] aims to increase their hydrogen storage capacity. For example, Li-doped carbon nanotubes [157] have shown improved hydrogen uptake due to the increased interaction between hydrogen molecules and the doped sites. However, the resource-intensive production of advanced carbon materials and their end-of-life disposal remain sustainability concerns.

MOFs represent a novel class of hydrogen storage materials that have gained significant attention in recent years, which composed of metal ions coordinated to organic ligands, forming a highly porous crystalline structure. This porosity allows for a high surface area, making MOFs excellent candidates for hydrogen physisorption. The tunability of MOFs, where both the metal centers and organic linkers can be modified, offers the potential to optimize their hydrogen storage properties [161]. For example, MOF-5 [162] and Hong Kong University of Science and Technology-1 (HKUST-1, known as MOF-199 or Cu3(BTC)2 (BTC = benzene-1,3,5-tricarboxylate, Ph(COO-)3)) [163] have demonstrated moderate hydrogen storage capacities at cryogenic temperatures, and ongoing research is focused on designing MOFs with functionalized pores [164] that can improve hydrogen binding energies at higher temperatures. Despite their potential, challenges such as stability during cycling and the environmental impact of large-scale production remain barriers to widespread adoption. Efforts to recycle MOFs [165], such as recovering metal ions or reusing ligands, are being investigated to mitigate these issues.

Another promising material class is complex hydrides, such as borohydrides (e.g. LiBH4 [166, 167], NaBH4 [168]) and alanates (e.g. NaAlH4 [169]), offering high gravimetric hydrogen capacities (e.g. LiBH4 can store up to 18.5 wt.% hydrogen [170]), but their practical use is limited by the high desorption temperatures and slow kinetics of hydrogen release. Advances in catalysis and nanoconfinement [138] have shown potential for improving the hydrogen release properties of complex hydrides. For instance, Ti-doped NaAlH4 [171] has been shown to reduce the desorption temperature and improve the reversibility of the hydrogenation/dehydrogenation cycles. However, the chemical reactivity and potential toxicity of some complex hydrides raise concerns about environmental sustainability and disposal.

Beyond solid-state materials such as metal hydrides, carbon-based compounds, and MOFs, alternative storage solutions, including liquid organic hydrogen carriers (LOHCs) and e-fuels (synthetic fuels), have emerged as promising candidates for scalable and transportable hydrogen storage. LOHCs are organic compounds capable of chemically binding and releasing hydrogen through catalytic processes [172]. Examples include toluene/methylcyclohexane [173], dibenzyltoluene [174], and N-ethylcarbazole systems [175]. These compounds store hydrogen via a hydrogenation reaction, such as converting toluene to methylcyclohexane, which can later release hydrogen through dehydrogenation. This process allows hydrogen to be stored in a liquid phase, operating at ambient temperatures and pressures, making it safer and more cost-effective than compressed or liquefied hydrogen. However, the hydrogenation and dehydrogenation processes require efficient catalysts and energy input, posing challenges for practical application [172]. Additionally, the stability of LOHCs over multiple cycles and the environmental impact of spills or degradation products require further research [176].

E-fuels store hydrogen chemically in the form of hydrocarbons or other energy carriers and are synthesized by combining green hydrogen with carbon dioxide or nitrogen. Common e-fuels include methane, ammonia, methanol, and hydrocarbons such as petrol, diesel, and kerosene [177]. Methane [178], produced through the Sabatier process by reacting hydrogen with CO2, serves as a direct substitute for natural gas and can be easily stored and integrated into existing gas grids. Ammonia [179], synthesized via the Haber-Bosch process, offers carbon-free hydrogen storage with an established global transport network, though its toxicity and energy-intensive production are challenges. Methanol [180], formed by reacting hydrogen with CO2, is a versatile liquid fuel used in engines and as a chemical precursor, and is also being developed for direct methanol fuel cells. Hydrocarbons [181] such as petrol, diesel, and kerosene, synthesized using green hydrogen and CO2 through the Fischer–Tropsch process, mimic conventional fossil fuels, enabling their use in existing transportation and industrial systems. While these fuels offer high energy densities and compatibility with current infrastructure, their production processes are energy-intensive, requiring renewable feedstocks and efficient catalysts to minimize environmental impact.

3.2 Thermodynamic and kinetic mechanisms of hydrogen storage

The performance of hydrogen storage materials is determined by the thermodynamics and kinetics of hydrogen adsorption and desorption processes. An ideal hydrogen storage material must strike a balance between high storage capacity, low operating temperatures, and fast kinetics for hydrogen uptake and release [137]. From a thermodynamic perspective, the enthalpy of hydrogen adsorption (ΔH) plays a critical role in determining the operating temperature of the storage system. For example, metal hydrides with strong metal-hydrogen bonds (high ΔH) typically require higher temperatures for hydrogen release, as the stored hydrogen is more tightly bound to the host material. Conversely, materials with weak hydrogen binding (low ΔH), such as physisorption-based materials, may release hydrogen at lower temperatures but suffer from lower storage capacities [182]. The trade-off between storage capacity and release temperature is one of the key challenges in hydrogen storage. For instance, light metal hydrides such as LiBH4 [166] offer very high gravimetric hydrogen densities but require temperatures exceeding 400°C for hydrogen release. On the other hand, materials like MOFs and carbon nanotubes can release hydrogen at near-ambient conditions but typically store less hydrogen per unit mass [162]. The kinetics of hydrogen storage materials are equally important. Slow adsorption/desorption kinetics can limit the rate at which hydrogen is stored or released, making the material unsuitable for real-time applications. Metal hydrides, in particular, often exhibit sluggish kinetics due to slow hydrogen diffusion through the solid lattice [183]. To overcome this, as discussed above, researchers are exploring the use of catalysts to accelerate the hydrogenation and dehydrogenation processes, and additionally, nanostructuring metal hydrides can shorten diffusion pathways, improving reaction kinetics and reducing desorption temperatures.

3.3 Applications and operational performance

In practical applications, hydrogen storage materials must not only demonstrate high performance in controlled laboratory conditions but also perform reliably under operational conditions, including factors such as cycling stability, thermal management, and material cost [137]. Cycling stability is critical for long-term use, as hydrogen storage materials must withstand multiple charge/discharge cycles without significant degradation. For example, metal hydrides can suffer from material pulverization and sintering after repeated hydrogenation/dehydrogenation cycles, which reduces their surface area and hydrogen capacity [184]. To mitigate these issues, materials are being designed with protective coatings [185] or nanostructured materials [147, 186] that maintain the integrity of the storage medium over time. Thermal management is another important consideration, particularly for high-temperature hydrides. Effective heat exchange systems must be incorporated into hydrogen storage devices to efficiently manage the exothermic and endothermic nature of hydrogen adsorption and desorption [187]. Advances in thermal conductivity enhancers, such as incorporating graphene [154] or metal hydride [188, 189] into storage materials, have shown promise in improving heat dissipation and reducing the energy input required for hydrogen release. Lastly, material cost and scalability are crucial for real-world deployment. While materials like MOFs and nanostructured metal hydrides show excellent performance in research settings, their high production costs and complexity of synthesis currently hinder large-scale adoption [136]. For hydrogen storage technologies to be commercially feasible, materials must be not only efficient but also cost-effective and easy to produce at scale, which is especially important for industrial and transportation applications, where the cost per unit of stored energy is a significant factor [190]. Metal hydrides, for instance, while efficient, often involve expensive materials or catalysts, and their energy-intensive production processes further add to the cost [187]. On the other hand, carbon-based materials, such as activated carbon or carbon nanotubes, are relatively more cost-effective, but their lower hydrogen storage capacities present a trade-off [156]. Advances in scalable production methods, such as template synthesis for MOFs [191] or ball-milling techniques for hydrides [192], are being explored to reduce costs while maintaining high performance. In practical applications, the choice of hydrogen storage material will depend on the specific requirements of the system. For stationary applications, where weight and volume constraints are less critical, metal hydrides with high volumetric capacity and stable cycling performance may be preferred [132]. In contrast, for mobile applications such as hydrogen-powered vehicles, materials with lower weight, faster kinetics, and the ability to operate at near-ambient conditions, such as lightweight carbon materials or advanced chemical hydrides, may be more suitable [193].

4. Hybrid systems for solar hydrogen energy applications

As the transition towards cleaner energy systems, it is crucial to explore how solar hydrogen technologies can be effectively integrated with existing renewable energy sources, storage solutions, and energy distribution systems. The effective integration of solar hydrogen production with PV, thermal energy, and battery storage technologies can enhance overall system efficiency, enable better energy management, and contribute to grid stability [194]. This section discusses the scientific and technical challenges of integrating solar hydrogen with other technologies and highlights potential solutions for optimizing these hybrid systems, as illustrated in Fig. 8, which provides an overview of hybrid systems for solar hydrogen energy applications.

A diagram outlining hybrid systems for solar hydrogen energy applications, categorized into: Integration with Renewable Energy Systems: Includes PV systems, CSP systems, and hybrid PV-thermal systems, focusing on energy flow optimization and thermal management. Coupling with Energy Storage: Covers battery storage and hydrogen storage integrated with fuel cells, enabling long-term energy use. Smart Grids and Energy Management: Discusses smart grids and AI-driven optimization for efficient energy distribution and cost savings. Case Studies and Pilot Projects: Highlights projects like SOLAR-JET, HyPSTER, and FH2R, demonstrating scalable applications and associated challenges. Technological and System Challenges: Addresses energy flow optimization, minimizing energy losses, and ensuring grid stability. Future Directions: Explores renewable energy hubs and advancements in power electronics for system improvement.
Figure 8.

An illustration of hybrid systems for solar hydrogen energy applications.

4.1 Advanced integration of solar hydrogen with energy systems

Integrating solar hydrogen production with PV or thermal energy systems presents several scientific and engineering challenges, in which solar hydrogen production requires efficient energy conversion from sunlight, and both PV and CSP systems can serve as primary energy sources for hydrogen production [194]. However, balancing the energy flows between electricity generation (via PV or CSP) and hydrogen production presents a significant challenge, particularly in terms of managing energy availability, storage, and conversion losses [195]. In a typical integrated system, PV cells or solar thermal collectors convert sunlight into electricity, which is then used to power electrolysis units that split water into hydrogen and oxygen. The efficiency of this process depends on several factors [196], including the capacity of the PV or CSP systems, the efficiency of the electrolysis units, and the availability of solar energy. A major challenge in this integration is managing the intermittency of solar energy, which can cause fluctuations in the power supply and affect the continuity of hydrogen production. Moreover, the design of hybrid solar systems that combine PV electricity generation with solar thermal hydrogen production (e.g. thermochemical cycles) must account for energy flow optimization. One emerging strategy [197] is the use of dual-use solar collectors that can simultaneously generate electricity and heat for thermochemical hydrogen production, requiring advanced thermal management and energy storage solutions to ensure that heat generated by the CSP system can be stored and used during periods of low solar radiation. Hybrid energy systems that combine solar hydrogen production with battery storage or fuel cells offer an additional level of flexibility. Batteries can be used to store excess electricity generated by PV systems, which can then be used to power electrolysis units during periods of low solar availability [198]. Alternatively, hydrogen produced during periods of excess solar energy can be stored and used to generate electricity through fuel cells during times of peak demand or when solar energy is unavailable. These hybrid systems enable energy shifting, where energy generated during periods of high solar availability is stored and used later, enhancing grid stability and reducing reliance on fossil fuel-based backup power systems [199].

4.1.1 Challenges in hybrid systems

One of the primary challenges in integrating solar hydrogen production with other renewable energy systems is the optimization of energy flows. In hybrid systems, there is a constant need to balance the supply and demand of energy between electricity generation, hydrogen production, and storage. Effective management of energy flow requires the integration of smart energy management systems that can dynamically distribute energy between PV systems, electrolysis units, batteries, and fuel cells, accounting for real-time variations [200] in solar irradiance, energy demand, and storage capacity to minimize energy losses and maximize efficiency. Advances in power electronics, such as bidirectional inverters [201] and intelligent controllers [202], play a crucial role in this optimization. For instance, by incorporating real-time data on solar irradiance, temperature, and grid demand, power electronics can regulate the power supply to electrolysis units or divert excess energy to battery storage, improving overall system performance [198]. Additionally, artificial intelligence (AI)-based algorithms [203] are being explored to predict energy demand and optimize the distribution of energy between hydrogen production and storage systems.

Integrating solar hydrogen into energy systems demands a comprehensive analysis of strategies to enhance system-level efficiency. In hybrid systems, energy losses can occur at several points [204], including electrolysis, hydrogen compression/storage, and conversion back to electricity via fuel cells. Reducing these losses is critical to improving the overall efficiency of the system. One approach to improve efficiency is to develop high-efficiency electrolyzers, such as proton exchange membrane [205, 206] or solid oxide electrolyzers [207] that can operate at lower energy input levels. In addition, optimizing thermal coupling between solar thermal systems and hydrogen production processes, such as using waste heat from CSP plants to preheat the water for electrolysis [208], can further reduce energy consumption.

Another scientific challenge in the integration of solar hydrogen systems with other technologies is optimizing the interaction between hydrogen production, energy storage, and grid stability in integrated systems. The intermittent nature of solar energy complicates efforts to maintain grid stability, particularly when large amounts of renewable energy are fed into the grid [209]. Hydrogen storage offers a potential solution by acting as a long-term storage medium that can absorb excess energy during periods of high solar generation and release energy during periods of low generation. However, the challenge lies in ensuring that hydrogen production and consumption are properly coordinated with grid demand. Integrating hydrogen-based systems with the grid requires the management of dispatchable power sources. Hydrogen stored during peak solar periods can be converted back into electricity through fuel cells to provide grid-balancing services. Fuel cells offer a clean and flexible means of generating electricity on demand, helping to stabilize the grid and compensate for fluctuations in solar power generation [210]. However, to achieve this level of integration, advanced grid infrastructure capable of accommodating variable hydrogen production and dispatchable hydrogen-powered electricity generation is required.

4.1.2 Smart grids and artificial intelligence-based optimization systems

The integration of solar hydrogen systems with renewable energy technologies hinges on advanced energy management solutions, with smart grids and AI-based optimization systems playing a pivotal role in enhancing efficiency, reliability, and flexibility. Smart grids, as intelligent energy distribution networks, dynamically manage the flow of energy between generation, storage, and consumption using real-time data and advanced communication systems [211]. In solar hydrogen systems, smart grids ensure surplus solar electricity is allocated to electrolysis units for hydrogen production during periods of high solar availability, while stored hydrogen can be converted back to electricity through fuel cells during low solar irradiance or high energy demand [202]. This real-time energy distribution reduces waste, enhances system efficiency, and contributes to grid stability.

AI-based optimization systems further complement smart grids by enabling predictive and adaptive energy management. By processing extensive datasets, including solar irradiance, temperature, energy demand, and grid load, AI algorithms forecast energy production and consumption patterns [212]. These insights allow for proactive adjustments, such as scheduling hydrogen production during peak solar generation or deploying stored hydrogen during periods of high demand. Additionally, AI optimizes the operation of electrolyzers by adjusting parameters like voltage and current to maximize hydrogen production efficiency, while managing thermal flows in systems combining solar thermal and hydrogen production to ensure effective use of CSP-generated heat [213].

AI-driven smart grids also play a crucial role in maintaining grid stability by monitoring grid performance, detecting potential instabilities from renewable energy fluctuations, and coordinating hydrogen storage systems and other renewable sources. Machine learning models integrated within these systems continuously enhance performance by analyzing historical data, enabling precise timing of hydrogen production to reduce costs and ensure reliability. Furthermore, AI-based demand response mechanisms allow smart grids to interact with end-users, promoting energy consumption behaviors that align with renewable energy availability [203].

The coupling of smart grids and AI in hybrid systems fosters a holistic energy management approach, enabling energy produced by solar hydrogen systems to be efficiently shared across transportation, industrial, and residential sectors. This sector coupling not only enhances system resilience but also accelerates the decarbonization of multiple energy-intensive sectors [214]. Expanding the role of smart grids and AI-based optimization systems is essential for overcoming challenges such as energy intermittency, resource allocation, and grid integration, ensuring that solar hydrogen technologies are seamlessly incorporated into a sustainable and adaptive energy infrastructure capable of meeting the demands of a low-carbon future.

4.2 Case studies and pilot projects

Case studies and pilot projects that have successfully demonstrated solar hydrogen production and its integration with other technologies provide valuable practical insights and underscore the feasibility of these systems. Table 5 presents an overview of case studies and pilot projects, highlighting their technological focus, efficiency, advantages, and associated challenges. For example, the Solar chemical reactor demonstration and Optimization for Long-term Availability of Renewable JET fuel (SOLAR-JET) project, funded by the European Union, successfully demonstrated efficiencies of up to 15%–20% in converting solar energy to hydrogen and synthetic fuels. While the approach is promising for industrial-scale hydrogen production, the primary challenges include the high temperatures (>1500°C) required for the reactions and the cost of CSP infrastructure [215, 216]. The integration of solar thermal energy with chemical processes proved viable but emphasized the need for advanced materials to withstand extreme thermal conditions and reduce operational costs.

Table 5.

Case studies and pilot projects: focus, efficiency, advantages, and challenges

ProjectTechnology FocusEfficiency (%)AdvantagesChallengesCitations
SOLAR-JETCSP with Thermochemical Cycles15–20High scalability potential for industrial useHigh temperatures and costly CSP infrastructure[215, 216]
HyPSTERPV Electrolysis + Salt Cavern Storage60–70Scalable storage, grid stabilizationGeological dependence, compression infrastructure[217, 218]
HyMARCHybrid Dual-Purpose Collectors50–60Combined thermal and electrical utilizationComplexity in energy flow management[219]
FH2RLarge-Scale PV Electrolysis60–80Grid stability, diverse applicationsHigh costs of large-scale electrolyzers[220]
ProjectTechnology FocusEfficiency (%)AdvantagesChallengesCitations
SOLAR-JETCSP with Thermochemical Cycles15–20High scalability potential for industrial useHigh temperatures and costly CSP infrastructure[215, 216]
HyPSTERPV Electrolysis + Salt Cavern Storage60–70Scalable storage, grid stabilizationGeological dependence, compression infrastructure[217, 218]
HyMARCHybrid Dual-Purpose Collectors50–60Combined thermal and electrical utilizationComplexity in energy flow management[219]
FH2RLarge-Scale PV Electrolysis60–80Grid stability, diverse applicationsHigh costs of large-scale electrolyzers[220]
Table 5.

Case studies and pilot projects: focus, efficiency, advantages, and challenges

ProjectTechnology FocusEfficiency (%)AdvantagesChallengesCitations
SOLAR-JETCSP with Thermochemical Cycles15–20High scalability potential for industrial useHigh temperatures and costly CSP infrastructure[215, 216]
HyPSTERPV Electrolysis + Salt Cavern Storage60–70Scalable storage, grid stabilizationGeological dependence, compression infrastructure[217, 218]
HyMARCHybrid Dual-Purpose Collectors50–60Combined thermal and electrical utilizationComplexity in energy flow management[219]
FH2RLarge-Scale PV Electrolysis60–80Grid stability, diverse applicationsHigh costs of large-scale electrolyzers[220]
ProjectTechnology FocusEfficiency (%)AdvantagesChallengesCitations
SOLAR-JETCSP with Thermochemical Cycles15–20High scalability potential for industrial useHigh temperatures and costly CSP infrastructure[215, 216]
HyPSTERPV Electrolysis + Salt Cavern Storage60–70Scalable storage, grid stabilizationGeological dependence, compression infrastructure[217, 218]
HyMARCHybrid Dual-Purpose Collectors50–60Combined thermal and electrical utilizationComplexity in energy flow management[219]
FH2RLarge-Scale PV Electrolysis60–80Grid stability, diverse applicationsHigh costs of large-scale electrolyzers[220]

Another notable example is the Hydrogen Pilot STorage for large Ecosystem Replication (HyPSTER) project in Europe, which focuses on integrating green hydrogen production with underground storage and renewable energy sources like wind and solar. HyPSTER combines PV-driven electrolysis with underground salt cavern storage. Electrolyzer efficiency in the project reaches ~60%–70%, with hydrogen stored at high pressures for grid stabilization during peak demand [217, 218]. The ability to scale storage capacity is a significant advantage, making it suitable for balancing intermittent renewable energy sources like wind and solar. However, the need for suitable geological formations for salt caverns and the costs of compression and transport infrastructure limit its broader applicability. The HyPSTER project highlights the practicality of coupling renewable hydrogen production with long-term storage but underscores the need for cost-effective solutions for hydrogen retrieval and distribution.

In the United States, the U.S. Department of Energy (DOE)-supported Advanced Water Splitting Materials Consortium (Hydrogen Materials Advanced Research Consortium, HyMARC) has been conducting pilot studies to integrate hydrogen production with renewable energy systems. For instance, these systems achieve efficiencies of 50%–60% in overall energy utilization by combining PV electricity generation with thermal energy recovery for thermochemical hydrogen production [219]. While this approach improves system efficiency, the complexity of managing thermal and electrical energy flows adds to operational challenges. Advanced materials for electrolyzers and thermal storage improve durability and performance, but high initial capital costs remain a barrier for widespread deployment.

Additionally, Japan’s Fukushima Hydrogen Energy Research Field (FH2R) is a large-scale demonstration project integrating solar hydrogen production with grid services. FH2R operates 20 MW PV array powering electrolysis units with an efficiency of ~60%–80% under optimal conditions [220]. The project demonstrates the feasibility of producing hydrogen at a large scale for diverse applications, including fuel cells and industrial processes. Hydrogen stored during peak solar availability is used to stabilize the grid and provide dispatchable power. Key challenges include the intermittent nature of solar energy and the high costs associated with large-scale electrolyzers. Despite these challenges, FH2R has proven to be a scalable model for integrating solar hydrogen production with grid services, achieving cost reductions through economies of scale.

These projects collectively highlight the strengths and limitations of solar hydrogen technologies. While CSP systems like SOLAR-JET excel in industrial applications, their scalability is constrained by high temperatures and costs. PV-driven systems such as FH2R and HyPSTER are more adaptable to grid integration and balancing, offering practical solutions for renewable energy storage. HyMARC demonstrates the potential of hybrid approaches, combining electrical and thermal energy for higher system efficiencies. The comparative analysis underscores the importance of tailoring technology choices to specific operational and regional requirements to maximize the benefits of solar hydrogen systems.

4.3 Technological advances and future directions

Technological advances in energy storage, smart grids, and power electronics are crucial for the integration of solar hydrogen production with other energy systems. Battery systems are becoming increasingly efficient and cost-effective, providing short-term energy storage solutions that complement the long-term storage potential of hydrogen. A cost-benefit analysis must account for capital and operational costs, such as the high expense of electrolyzers and thermal management systems [221], against the long-term benefits of improved energy efficiency, reduced reliance on fossil fuels, and enhanced grid stability. Additionally, fuel cell technologies are advancing, offering higher efficiency and longer lifespans, making them suitable for a wide range of applications, from residential energy systems to large-scale power plants. One emerging trend is the development of integrated renewable energy hubs, where solar hydrogen production, battery storage, and fuel cells are co-located with other renewable energy sources such as wind and hydropower [222], allowing for energy-sharing between different technologies, improving overall system efficiency and reliability. Furthermore, advances in hydrogen transport and distribution infrastructure [223], such as hydrogen pipelines and liquid hydrogen storage, will enable the large-scale deployment of hydrogen as a key energy carrier in the renewable energy economy. Another area of innovation is the exploration of electrochemical hydrogen storage systems that combine elements of battery storage with hydrogen production. Technological advances are crucial for improving system integration and reducing costs. For instance, reversible fuel cells or flow batteries that can both store hydrogen and generate electricity are being researched as a means of simplifying the energy conversion process and improving overall system efficiency [224]. Another scientific challenge is optimizing the synergy between solar hydrogen production, energy storage, and grid stability [225], which requires the development of real-time control algorithms that can predict solar energy availability, hydrogen demand, and grid load in order to dynamically adjust the energy flow between different components of the system [226]. Additionally, the creation of grid-interactive hydrogen storage systems, where hydrogen production and fuel cell electricity generation can be used to smooth out fluctuations in renewable energy supply, will play a crucial role in ensuring grid stability [227], being seamlessly integrated into existing grid infrastructure, and ongoing research is focused on how to standardize and scale hydrogen infrastructure for use in large-scale energy systems and further enhance the economic viability of these systems.

5. Discussion and future perspectives

The development of solar hydrogen production and storage technologies presents a transformative opportunity to advance sustainable energy systems, yet their implementation faces significant technical and economic hurdles. Thermochemical, photochemical, and biological methods each offer distinct pathways for hydrogen production, but their scalability and viability remain subjects of ongoing debate, emphasizing the need for critical analysis and innovation.

Thermochemical hydrogen production, particularly through integration with CSP systems, has shown considerable promise due to its high efficiency and compatibility with advanced materials such as metal oxides in redox cycles. However, the reliance on extreme operating temperatures (>1500°C) presents challenges, including material degradation, complex reactor design, and high operational costs. While advocates highlight the potential of high-temperature materials and advanced thermal energy storage systems to mitigate these issues, concerns persist regarding the energy losses during heat transfer and the scalability of CSP infrastructure. Continued research into novel reactor designs and improved thermal coupling strategies is necessary to address these limitations effectively.

Photochemical hydrogen production provides a more direct and potentially cost-effective approach by harnessing sunlight to drive water-splitting reactions. Yet, the low solar-to-hydrogen conversion efficiencies of current photocatalysts represent a significant barrier. Advances in bandgap engineering, heterostructures, and plasmonic enhancements are promising, but questions remain about the long-term durability and stability of materials such as perovskites MOFs. While these materials show enhanced light absorption and charge separation capabilities, their environmental sensitivity, particularly to moisture, raises concerns about their practicality. Addressing these challenges will require robust testing under real-world conditions and the development of stable material configurations.

Biological hydrogen production is often lauded for its low-temperature processes and alignment with natural metabolic pathways, yet it struggles with slow production rates and the oxygen sensitivity of key enzymes such as hydrogenase. Innovations in metabolic engineering, including the development of oxygen-tolerant enzymes and synthetic biology techniques, have yielded promising results in laboratory settings. However, scaling these systems introduces additional complexities, such as nutrient management and efficient bioreactor design. Opposing perspectives on the feasibility of large-scale deployment highlight the necessity for pilot-scale projects that bridge the gap between experimental advancements and industrial applications.

Hydrogen storage technologies remain a pivotal factor in enabling the widespread adoption of solar hydrogen systems. Conventional methods like compressed gas and liquefied hydrogen are mature but constrained by low volumetric densities and energy-intensive processes. Solid-state storage materials, including metal hydrides and MOFs, offer safer and higher-density storage solutions, though they are challenged by slow hydrogen release kinetics, high material costs, and degradation during repeated cycling. Advocates for these technologies point to advancements in nanostructuring and catalytic enhancements as pathways to improve performance, while critics emphasize the need for comprehensive lifecycle analyses to ensure their cost-effectiveness and scalability.

Integrating solar hydrogen systems into renewable energy grids adds another layer of complexity. Hybrid systems that combine PV or CSP technologies with hydrogen production offer flexibility but require sophisticated energy management to balance supply and demand. While some view the energy losses in electrolyzers and fuel cells as a barrier to efficiency, advancements in smart grids and AI-driven optimization are being explored to dynamically allocate energy and enhance grid stability. However, the limited number of large-scale implementations highlights the uncertainty surrounding the scalability and reliability of such systems, underscoring the importance of demonstration projects to validate their potential.

The successful adoption of solar hydrogen technologies will depend not only on technological breakthroughs but also on supportive policy measures and economic incentives. Subsidies, tax credits, and carbon pricing mechanisms are essential to close the cost gap between renewable and fossil fuel-derived hydrogen. Furthermore, the establishment of international hydrogen certification schemes will promote transparency and distinguish sustainable hydrogen in global markets.

In conclusion, while solar hydrogen technologies hold significant potential for decarbonization and energy storage, addressing their technical, economic, and policy-related challenges will require a balanced approach that incorporates diverse perspectives. By fostering collaboration across research, industry, and government sectors, and prioritizing real-world testing, these technologies can evolve into a cornerstone of the global energy transition, supporting applications ranging from grid stabilization to industrial processes and sustainable transportation.

Author contributions

Ge Chen (Conceptualization [lead], Data curation [equal], Formal analysis [equal], Investigation [lead], Methodology [equal], Resources [equal], Software [equal], Validation [equal], Visualization [equal], Writing—original draft [lead]), Renhui Sun (Conceptualization [equal], Investigation [supporting], Resources [supporting], Validation [supporting]), and Baodong Wang (Conceptualization [equal], Formal analysis [equal], Investigation [equal], Methodology [equal], Resources [equal], Supervision [lead], Validation [equal], Writing—review & editing [lead])

Conflict of interest

None declared.

Funding

None declared.

Data availability

The data underlying this article is available in the article.

References

[1]

Abbass
K
,
Qasim
MZ
,
Song
H
et al.
A review of the global climate change impacts, adaptation, and sustainable mitigation measures
.
Environ Sci Pollut Res
2022
;
29
:
42539
59
. https://doi.org/

[2]

Wang
J
,
Azam
W.
Natural resource scarcity, fossil fuel energy consumption, and total greenhouse gas emissions in top emitting countries
.
Geosci Front
2024
;
15
:
101757
. https://doi.org/

[3]

Maka
AOM
,
Alabid
JM.
Solar energy technology and its roles in sustainable development
.
Clean Energy
2022
;
6
:
476
83
. https://doi.org/

[4]

Mitali
J
,
Dhinakaran
S
,
Mohamad
AA.
Energy storage systems: a review
.
Energy Storage Saving
2022
;
1
:
166
216
. https://doi.org/

[5]

Hren
R
,
Vujanović
A
,
van Fan
Y.
et al.
Hydrogen production, storage and transport for renewable energy and chemicals: an environmental footprint assessment
.
Renew Sustain Energy Rev
2023
;
173
:
113113
. https://doi.org/

[6]

Hassan
Q
,
Tabar
VS
,
Sameen
AZ
et al.
A review of green hydrogen production based on solar energy; techniques and methods
.
Energy Harvest Syst
2024
;
11
:
20220134
.  https://doi.org/

[7]

Nijsse
FJMM
,
Mercure
J-F
,
Ameli
N
et al.
The momentum of the solar energy transition
.
Nat Commun
2023
;
14
:
6542
. https://doi.org/

[8]

Sadeq
AM
,
Homod
RZ
,
Hussein
AK
et al.
Hydrogen energy systems: technologies, trends, and future prospects
.
Sci Total Environ
2024
;
939
:
173622
. https://doi.org/

[9]

Bian
H
,
Li
D
,
Yan
J
et al.
Perovskite – a wonder catalyst for solar hydrogen production
.
J Energy Chem
2021
;
57
:
325
40
. https://doi.org/

[10]

Li
W
,
Duan
W
,
Liao
G
et al.
0.68% of solar-to-hydrogen efficiency and high photostability of organic-inorganic membrane catalyst
.
Nat Commun
2024
;
15
:
6763
. https://doi.org/

[11]

Qureshy
AM
,
Dincer
I.
Development of a new solar photoelectrochemical reactor design for more efficient hydrogen production
.
Energy Convers Manage
2021
;
228
:
113714
. https://doi.org/

[12]

Murmura
MA
,
Annesini
MC.
Methodologies for the design of solar receiver/reactors for thermochemical hydrogen production
.
Processes
2020
;
8
:
308
. https://doi.org/

[13]

Messini
EMB
,
Bourek
Y
,
Ammari
C
et al.
The integration of solar-hydrogen hybrid renewable energy systems in oil and gas industries for energy efficiency: optimal sizing using Fick’s law optimisation algorithm
.
Energy Convers Manage
2024
;
308
:
118372
. https://doi.org/

[14]

Joshua
SR
,
Yeon
AN
,
Park
S
et al.
Solar–hydrogen storage system: architecture and integration design of university energy management systems
.
Appl Sci
2024
;
14
:
4376
. https://doi.org/

[15]

Das
D
,
Chakraborty
I
,
Bohre
AK
et al.
Sustainable integration of green hydrogen in renewable energy systems for residential and EV applications
.
Int J Energy Res
2024
;
2024
:
1
20
. https://doi.org/

[16]

Yilmaz
F
,
Selbaş
R.
Thermodynamic performance assessment of solar based sulfur-iodine thermochemical cycle for hydrogen generation
.
Energy
2017
;
140
:
520
9
. https://doi.org/

[17]

Zeng
J
,
Zhang
J
,
Ling
B
et al.
Hydrogen production by sulfur-iodine thermochemical cycle — current status and recent advances
.
Int J Hydrog Energy
2024
;
86
:
677
702
. https://doi.org/

[18]

Almeida Pazmiño
GA
,
Jung
S
,
Roh
S-H.
Process modeling of a hybrid-sulfur thermochemical cycle combined with solid oxide fuel cell/gas turbine system
.
Energy Convers Manage
2022
;
262
:
115669
. https://doi.org/

[19]

Gorensek
MB
,
Corgnale
C
,
Summers
WA.
Development of the hybrid sulfur cycle for use with concentrated solar heat. I. Conceptual design
.
Int J Hydrog Energy
2017
;
42
:
20939
54
. https://doi.org/

[20]

Yilmaz
F
,
Selbaş
R.
Investigation and performance assessment of calcium bromine cycle for hydrogen production
.
Int J Glob Warm
2019
;
17
:
279
. https://doi.org/

[21]

Mao
Y
,
Gao
Y
,
Dong
W
et al.
Hydrogen production via a two-step water splitting thermochemical cycle based on metal oxide – a review
.
Appl Energy
2020
;
267
:
114860
. https://doi.org/

[22]

Abanades
S.
Metal oxides applied to thermochemical water-splitting for hydrogen production using concentrated solar energy
.
ChemEngineering
2019
;
3
:
63
. https://doi.org/

[23]

Miller
JE
,
Allendorf
MD
,
Diver
RB
et al.
Metal oxide composites and structures for ultra-high temperature solar thermochemical cycles
.
J Mater Sci
2008
;
43
:
4714
28
. https://doi.org/

[24]

Ishaq
M
,
Dincer
I.
A cradle-to-gate life cycle assessment for clean hydrogen gas production pathway using the CeO2/Ce2O3-based redox thermochemical cycle
.
Gas Sci Eng
2024
;
131
:
205464
. https://doi.org/

[25]

Rakitskaya
TL
,
Truba
AS
,
Ennan
AA
et al.
Synthesis and catalytic properties of iron oxides in the reaction of low-temperature ozone decomposition
.
Acta Phys Pol A
2018
;
133
:
1079
83
. https://doi.org/

[26]

Kapuścik
P
,
Wojcieszak
D
,
Pokora
P
et al.
Low temperature hydrogen sensor with high sensitivity based on CeOx thin film
.
Sens Actuators B
2024
;
417
:
136148
. https://doi.org/

[27]

Rukini
A
,
Rhamdhani
MA
,
Brooks
GA
et al.
Metals production and metal oxides reduction using hydrogen: a review
.
J Sustain Metall
2022
;
8
:
1
24
. https://doi.org/

[28]

Chen
C
,
Jiao
F
,
Lu
B
et al.
Challenges and perspectives for solar fuel production from water/carbon dioxide with thermochemical cycles
.
Carb Neutrality
2023
;
2
:
1
27
. https://doi.org/

[29]

Shanmugam
V
,
Neuberg
S
,
Zapf
R
et al.
Hydrogen production over highly active Pt based catalyst coatings by steam reforming of methanol: effect of support and co-support
.
Int J Hydrog Energy
2020
;
45
:
1658
70
. https://doi.org/

[30]

Zhang
G
,
Huang
X
,
Ma
X
et al.
A fast and general approach to produce a carbon coated Janus metal/oxide hybrid for catalytic water splitting
.
J Mater Chem A
2021
;
9
:
7606
16
. https://doi.org/

[31]

Deng
Z-Y
,
Zhu
L-L
,
Tang
Y-B
et al.
Role of particle sizes in hydrogen generation by the reaction of Al with water
.
J Am Ceram Soc
2010
;
93
:
2998
3001
. https://doi.org/

[32]

Izurieta
EM
,
Pedernera
MN
,
López
E.
Study of a thermally integrated parallel plates reactor for hydrogen production
.
Chem Eng Sci
2019
;
196
:
344
53
. https://doi.org/

[33]

Meng
Q-L
,
Lee
C
,
Ishihara
T
et al.
Reactivity of CeO2-based ceramics for solar hydrogen production via a two-step water-splitting cycle with concentrated solar energy
.
Int J Hydrog Energy
2011
;
36
:
13435
41
. https://doi.org/

[34]

Wong
B
,
Buckingham
RT
,
Brown
LC
et al.
Construction materials development in sulfur–iodine thermochemical water-splitting process for hydrogen production
.
Int J Hydrog Energy
2007
;
32
:
497
504
. https://doi.org/

[35]

Neises
M
,
Roeb
M
,
Schmücker
M
et al.
Kinetic investigations of the hydrogen production step of a thermochemical cycle using mixed iron oxides coated on ceramic substrates
.
Int J Energy Res
2010
;
34
:
651
61
. https://doi.org/

[36]

Hildebrandt
AF
,
Rose
KA.
Receiver design considerations for solar central receiver hydrogen production
.
Sol Energy
1985
;
35
:
199
206
. https://doi.org/

[37]

Ma
Z
,
Davenport
P
,
Saur
G.
System and technoeconomic analysis of solar thermochemical hydrogen production
.
Renew Energy
2022
;
190
:
294
308
. https://doi.org/

[38]

Ehrhart
BD
,
Muhich
CL
,
Al-Shankiti
I
et al.
System efficiency for two-step metal oxide solar thermochemical hydrogen production – part 2: impact of gas heat recuperation and separation temperatures
.
Int J Hydrog Energy
2016
;
41
:
19894
903
. https://doi.org/

[39]

Singla
S
,
Shetti
NP
,
Basu
S
et al.
Hydrogen production technologies – membrane based separation, storage and challenges
.
J Environ Manage
2022
;
302
:
113963
. https://doi.org/

[40]

David
E
,
Kopac
J.
Devlopment of palladium/ceramic membranes for hydrogen separation
.
Int J Hydrog Energy
2011
;
36
:
4498
506
. https://doi.org/

[41]

Liu
C
,
Park
J
,
de Santiago
HA
et al.
Perovskite oxide materials for solar thermochemical hydrogen production from water splitting through chemical looping
.
ACS Catal
2024
;
14
:
14974
5013
. https://doi.org/

[42]

Müller
R
,
Haeberling
P
,
Palumbo
RD.
Further advances toward the development of a direct heating solar thermal chemical reactor for the thermal dissociation of ZnO(s)
.
Sol Energy
2006
;
80
:
500
11
. https://doi.org/

[43]

Onigbajumo
A
,
Swarnkar
P
,
Will
G
et al.
Techno-economic evaluation of solar-driven ceria thermochemical water-splitting for hydrogen production in a fluidized bed reactor
.
J Clean Prod
2022
;
371
:
133303
. https://doi.org/

[44]

Rahman
MA
,
Parvej
AM
,
Aziz
MA.
Concentrating technologies with reactor integration and effect of process variables on solar assisted pyrolysis: a critical review
.
Therm Sci Eng Prog
2021
;
25
:
100957
. https://doi.org/

[45]

Dutton
R.
Materials degradation problems in hydrogen energy systems☆
.
Int J Hydrog Energy
1984
;
9
:
147
55
. https://doi.org/

[46]

Yvon
P
,
Carré
F.
Structural materials challenges for advanced reactor systems
.
J Nucl Mater
2009
;
385
:
217
22
. https://doi.org/

[47]

Burulday
ME
,
Mert
MS
,
Javani
N.
Thermodynamic analysis of a parabolic trough solar power plant integrated with a biomass-based hydrogen production system
.
Int J Hydrog Energy
2022
;
47
:
19481
501
. https://doi.org/

[48]

Lee
JE
,
Shafiq
I
,
Hussain
M
et al.
A review on integrated thermochemical hydrogen production from water
.
Int J Hydrog Energy
2022
;
47
:
4346
56
. https://doi.org/

[49]

Ismail
AA
,
Bahnemann
DW.
Photochemical splitting of water for hydrogen production by photocatalysis: a review
.
Sol Energy Mater Sol Cells
2014
;
128
:
85
101
. https://doi.org/

[50]

Teets
TS
,
Nocera
DG.
Photocatalytic hydrogen production
.
Chem Commun
2011
;
47
:
9268
74
. https://doi.org/

[51]

Kitano
M
,
Tsujimaru
K
,
Anpo
M.
Hydrogen production using highly active titanium oxide-based photocatalysts
.
Top Catal
2008
;
49
:
4
17
. https://doi.org/

[52]

Ismael
M.
Enhanced photocatalytic hydrogen production and degradation of organic pollutants from Fe (III) doped TiO2 nanoparticles
.
J Environ Chem Eng
2020
;
8
:
103676
. https://doi.org/

[53]

Madhumitha
A
,
Preethi
V
,
Kanmani
S.
Photocatalytic hydrogen production using TiO2 coated iron-oxide core shell particles
.
Int J Hydrog Energy
2018
;
43
:
3946
56
. https://doi.org/

[54]

Pérez-Larios
A
,
Torres-Ramos
MI
,
González-Vargas
OA
et al.
Ti–Fe mixed oxides as photocatalysts in the generation of hydrogen under UV-light irradiation
.
Int J Hydrog Energy
2022
;
47
:
30178
86
. https://doi.org/

[55]

Shanmugaratnam
S
,
Velauthapillai
D
,
Ravirajan
P
et al.
CoS2/TiO2 nanocomposites for hydrogen production under UV irradiation
.
Materials
2019
;
12
:
3882
. https://doi.org/

[56]

Shim
E
,
Park
Y
,
Bae
S
et al.
Photocurrent by anodized TiO2 photoelectrode for enzymatic hydrogen production and chromium(VI) reduction
.
Int J Hydrog Energy
2008
;
33
:
5193
8
. https://doi.org/

[57]

Samokhvalov
A.
Hydrogen by photocatalysis with nitrogen codoped titanium dioxide
.
Renew Sustain Energy Rev
2017
;
72
:
981
1000
. https://doi.org/

[58]

Reverberi
AP
,
Klemeš
JJ
,
Varbanov
PS
et al.
A review on hydrogen production from hydrogen sulphide by chemical and photochemical methods
.
J Clean Prod
2016
;
136
:
72
80
. https://doi.org/

[59]

Yang
C
,
Wang
Z
,
Lin
T
et al.
Core-shell nanostructured “black” rutile titania as excellent catalyst for hydrogen production enhanced by sulfur doping
.
J Am Chem Soc
2013
;
135
:
17831
8
. https://doi.org/

[60]

Almomani
F
,
Shawaqfah
M
,
Alkasrawi
M.
Solar-driven hydrogen production from a water-splitting cycle based on carbon-TiO2 nano-tubes
.
Int J Hydrog Energy
2022
;
47
:
3294
305
. https://doi.org/

[61]

Zhao
Y
,
Hoivik
N
,
Wang
K.
Recent advance on engineering titanium dioxide nanotubes for photochemical and photoelectrochemical water splitting
.
Nano Energy
2016
;
30
:
728
44
. https://doi.org/

[62]

Sun
Y
,
Wang
G
,
Yan
K.
TiO2 nanotubes for hydrogen generation by photocatalytic water splitting in a two-compartment photoelectrochemical cell
.
Int J Hydrog Energy
2011
;
36
:
15502
8
. https://doi.org/

[63]

Zhang
Y
,
Shang
M
,
Mi
Y
et al.
Influence of the amount of hydrogen fluoride on the formation of (001)‐faceted titanium dioxide nanosheets and their photocatalytic hydrogen generation performance
.
ChemPlusChem
2014
;
79
:
1159
66
. https://doi.org/

[64]

Zeng
D
,
Yang
L
,
Zhou
P
et al.
Au Cu alloys deposited on titanium dioxide nanosheets for efficient photocatalytic hydrogen evolution
.
Int J Hydrog Energy
2018
;
43
:
15155
63
. https://doi.org/

[65]

Singh
AP
,
Arora
P
,
Basu
S
et al.
Graphitic carbon nitride based hydrogen treated disordered titanium dioxide core-shell nanocatalyst for enhanced photocatalytic and photoelectrochemical performance
.
Int J Hydrog Energy
2016
;
41
:
5617
28
. https://doi.org/

[66]

Kim
CY
,
Park
JH
,
Kim
TG.
Effect of photochemical hydrogen doping on the electrical properties of ZnO thin-film transistors
.
J Alloys Compd
2018
;
732
:
300
5
. https://doi.org/

[67]

Sampaio
MJ
,
Oliveira
JW
,
Sombrio
CI
et al.
Photocatalytic performance of Au/ZnO nanocatalysts for hydrogen production from ethanol
.
Appl Catal A
2016
;
518
:
198
205
. https://doi.org/

[68]

Knoks
A
,
Kleperis
J
,
Bajars
G
et al.
WO3 as additive for efficient photocatalyst binary system TiO2 /WO3
.
Latv J Phys Tech Sci
2021
;
58
:
24
34
. https://doi.org/

[69]

Tahir
AA
,
Wijayantha
KGU
,
Saremi-Yarahmadi
S
et al.
Nanostructured α-Fe2O3 thin films for photoelectrochemical hydrogen generation
.
Chem Mater
2009
;
21
:
3763
72
. https://doi.org/

[70]

Frites
M
,
Shaban
YA
,
Khan
SU.
Iron oxide (n-Fe2O3) nanowire films and carbon modified (CM)-n-Fe2O3 thin films for hydrogen production by photosplitting of water
.
Int J Hydrog Energy
2010
;
35
:
4944
8
. https://doi.org/

[71]

Guan
Z
,
Wang
P
,
Li
Q
et al.
Remarkable enhancement in solar hydrogen generation from MoS2 -RGO/ZnO composite photocatalyst by constructing a robust electron transport pathway
.
Chem Eng J
2017
;
327
:
397
405
. https://doi.org/

[72]

Nisar
A
,
Saeed
M
,
Muneer
M
et al.
Synthesis and characterization of ZnO decorated reduced graphene oxide (ZnO-rGO) and evaluation of its photocatalytic activity toward photodegradation of methylene blue
.
Environ Sci Pollut Res
2022
;
29
:
418
30
. https://doi.org/

[73]

Neena
D
,
Kondamareddy
KK
,
Humayun
M
et al.
Fabrication of ZnO/N-rGO composite as highly efficient visible-light photocatalyst for 2,4-DCP degradation and H2 evolution
.
Appl Surf Sci
2019
;
488
:
611
9
. https://doi.org/

[74]

Tasleem
S
,
Tahir
M.
Current trends in strategies to improve photocatalytic performance of perovskites materials for solar to hydrogen production
.
Renew Sustain Energy Rev
2020
;
132
:
110073
. https://doi.org/

[75]

Tasleem
S
,
Tahir
M.
Recent progress in structural development and band engineering of perovskites materials for photocatalytic solar hydrogen production: a review
.
Int J Hydrog Energy
2020
;
45
:
19078
111
. https://doi.org/

[76]

Getoff
N.
Basic problems of photochemical and photoelectrochemical hydrogen production from water
.
Int J Hydrog Energy
1984
;
9
:
997
1004
. https://doi.org/

[77]

Waller
MR
,
Townsend
TK
,
Zhao
J
et al.
Single-crystal tungsten oxide nanosheets: photochemical water oxidation in the quantum confinement regime
.
Chem Mater
2012
;
24
:
698
704
. https://doi.org/

[78]

Lee
CW
,
Lee
BH
,
Park
S
et al.
Photochemical tuning of dynamic defects for high-performance atomically dispersed catalysts
.
Nat Mater
2024
;
23
:
552
9
. https://doi.org/

[79]

Yao
W
,
Song
X
,
Huang
C
et al.
Enhancing solar hydrogen production via modified photochemical treatment of Pt/CdS photocatalyst
.
Catal Today
2013
;
199
:
42
47
. https://doi.org/

[80]

Parayil
SK
,
Kibombo
HS
,
Wu
C-M
et al.
Synthesis-dependent oxidation state of platinum on TiO2 and their influences on the solar simulated photocatalytic hydrogen production from water
.
J Phys Chem C
2013
;
117
:
16850
62
. https://doi.org/

[81]

Lei
J-M
,
Peng
Q-X
,
Luo
S-P
et al.
A nickel complex, an efficient cocatalyst for both electrochemical and photochemical driven hydrogen production from water
.
Mol Catal
2018
;
448
:
10
17
. https://doi.org/

[82]

Xu
Y
,
Xu
R.
Nickel-based cocatalysts for photocatalytic hydrogen production
.
Appl Surf Sci
2015
;
351
:
779
93
. https://doi.org/

[83]

Li
Y-Y
,
Wang
J-H
,
Luo
Z-J
et al.
Plasmon-enhanced photoelectrochemical current and hydrogen production of (MoS2-TiO2)/Au hybrids
.
Sci Rep
2017
;
7
:
7178
. https://doi.org/

[84]

Belessiotis
GV
,
Kontos
AG.
Plasmonic silver (Ag)-based photocatalysts for H2 production and CO2 conversion: review, analysis and perspectives
.
Renew Energy
2022
;
195
:
497
515
. https://doi.org/

[85]

Chen
L
,
Chen
R
,
Hu
H
et al.
Enhancement of photocatalytic hydrogen production of semiconductor by plasmonic silver nanocubes under visible light
.
Mater Lett
2019
;
242
:
47
50
. https://doi.org/

[86]

Thangamuthu
M
,
Vankayala
K
,
Xiong
L
et al.
Tungsten oxide-based Z-scheme for visible light-driven hydrogen production from water splitting
.
ACS Catal
2023
;
13
:
9113
24
. https://doi.org/

[87]

Li
J
,
Yuan
H
,
Zhang
W
et al.
Advances in Z‐scheme semiconductor photocatalysts for the photoelectrochemical applications: a review
.
Carbon Energy
2022
;
4
:
294
331
. https://doi.org/

[88]

Yang
Y
,
Jing
D
,
Zhao
L.
Computational fluid dynamics modeling of reactive multiphase flow for suspended photocatalytic water splitting of hydrogen production system
.
Appl Therm Eng
2020
;
173
:
115220
. https://doi.org/

[89]

Wang
G
,
Lv
S
,
Shen
Y
et al.
Advancements in heterojunction, cocatalyst, defect and morphology engineering of semiconductor oxide photocatalysts
.
J Materiomics
2024
;
10
:
315
38
. https://doi.org/

[90]

Khalid
NR
,
Ahmed
E
,
Hong
Z
et al.
Enhanced photocatalytic activity of graphene–TiO2 composite under visible light irradiation
.
Cur Appl Phys
2013
;
13
:
659
63
. https://doi.org/

[91]

Shen
M
,
Zhang
Y
,
Xu
H
et al.
MOFs based on the application and challenges of perovskite solar cells
.
iScience
2021
;
24
:
103069
. https://doi.org/

[92]

Hou
J
,
Wang
Z
,
Chen
P
et al.
Intermarriage of halide perovskites and metal‐organic framework crystals
.
Angew Chem
2020
;
132
:
19602
17
. https://doi.org/

[93]

Pullen
S
,
Ott
S.
Photochemical hydrogen production with metal–organic frameworks
.
Top Catal
2016
;
59
:
1712
21
. https://doi.org/

[94]

Zhang
S
,
Han
G.
Intrinsic and environmental stability issues of perovskite photovoltaics
.
Prog Energy
2020
;
2
:
22002
. https://doi.org/

[95]

Ren
M
,
Qian
X
,
Chen
Y
et al.
Potential lead toxicity and leakage issues on lead halide perovskite photovoltaics
.
J Hazard Mater
2022
;
426
:
127848
. https://doi.org/

[96]

Chen
B
,
Wang
S
,
Song
Y
et al.
A critical review on the moisture stability of halide perovskite films and solar cells
.
Chem Eng J
2022
;
430
:
132701
. https://doi.org/

[97]

Angelis
Fde.
The prospect of lead-free perovskite photovoltaics
.
ACS Energy Lett
2021
;
6
:
1586
7
. https://doi.org/

[98]

Xia
J
,
Liang
C
,
Gu
H
et al.
Surface passivation toward efficient and stable perovskite solar cells
.
Energy Environ Mater
2023
;
6
:
12296
.  https://doi.org/

[99]

Raman
RK
,
Gurusamy Thangavelu
SA
,
Venkataraj
S
et al.
Materials, methods and strategies for encapsulation of perovskite solar cells: from past to present
.
Renew Sustain Energy Rev
2021
;
151
:
111608
. https://doi.org/

[100]

Krishnamurthy
S
,
Pandey
P
,
Kaur
J
et al.
Organic–inorganic hybrid and inorganic halide perovskites: structural and chemical engineering, interfaces and optoelectronic properties
.
J Phys D Appl Phys
2021
;
54
:
133002
. https://doi.org/

[101]

Akhlaghi
N
,
Najafpour-Darzi
G.
A comprehensive review on biological hydrogen production
.
Int J Hydrog Energy
2020
;
45
:
22492
512
. https://doi.org/

[102]

Teke
GM
,
Anye Cho
B
,
Bosman
CE
et al.
Towards industrial biological hydrogen production: a review
.
World J Microbiol Biotechnol
2023
;
40
:
37
. https://doi.org/

[103]

Golding
A-L
,
Dong
Z.
Hydrogen production by nitrogenase as a potential crop rotation benefit
.
Environ Chem Lett
2010
;
8
:
101
21
. https://doi.org/

[104]

Xuan
J
,
He
L
,
Wen
W
et al.
Hydrogenase and nitrogenase: key catalysts in biohydrogen production
.
Molecules
2023
;
28
:
1392
. https://doi.org/

[105]

Benemann
JR.
Hydrogen production by microalgae
.
J Appl Phycol
2000
;
12
:
291
300
. https://doi.org/

[106]

Kamshybayeva
GK
,
Kossalbayev
BD
,
Sadvakasova
AK
et al.
Genetic engineering contribution to developing cyanobacteria-based hydrogen energy to reduce carbon emissions and establish a hydrogen economy
.
Int J Hydrog Energy
2024
;
54
:
491
511
. https://doi.org/

[107]

McKinlay
JB
,
Harwood
CS.
Calvin cycle flux, pathway constraints, and substrate oxidation state together determine the H2 biofuel yield in photoheterotrophic bacteria
.
mBio
2011
;
2
:
e00323
10
. https://doi.org/

[108]

Seaver
LC
,
Imlay
JA.
Are respiratory enzymes the primary sources of intracellular hydrogen peroxide
?.
J Biol Chem
2004
;
279
:
48742
50
. https://doi.org/

[109]

Abd-Elrahman
NK
,
Al-Harbi
N
,
Al-Hadeethi
Y
et al.
Influence of nanomaterials and other factors on biohydrogen production rates in microbial electrolysis cells-a review
.
Molecules
2022
;
27
:
8594
. https://doi.org/

[110]

Han
H-X
,
Tian
L-J
,
Liu
D-F
et al.
Reversing electron transfer chain for light-driven hydrogen production in biotic-abiotic hybrid systems
.
J Am Chem Soc
2022
;
144
:
6434
41
. https://doi.org/

[111]

Yildirim
O
,
Tunay
D
,
Ozkaya
B
et al.
Effect of green synthesized silver oxide nanoparticle on biological hydrogen production
.
Int J Hydrog Energy
2022
;
47
:
19517
25
. https://doi.org/

[112]

Leone
L
,
Sgueglia
G
,
La Gatta
S
et al.
Enzymatic and bioinspired systems for hydrogen production
.
Int J Mol Sci
2023
;
24
:
8605
. https://doi.org/

[113]

López-Calixto
CG
,
Barawi
M
,
Gomez-Mendoza
M
et al.
Hybrids based on BOPHY-conjugated porous polymers as photocatalysts for hydrogen production: insight into the charge transfer pathway
.
ACS Catal
2020
;
10
:
9804
12
. https://doi.org/

[114]

Dutta
D
,
De
D
,
Chaudhuri
S
et al.
Hydrogen production by Cyanobacteria
.
Microb Cell Fact
2005
;
4
:
36
. https://doi.org/

[115]

Guo
C-L
,
Zhu
X
,
Liao
Q
et al.
Enhancement of photo-hydrogen production in a biofilm photobioreactor using optical fiber with additional rough surface
.
Bioresour Technol
2011
;
102
:
8507
13
. https://doi.org/

[116]

Chen
C-Y
,
Lee
C-M
,
Chang
J-S.
Hydrogen production by indigenous photosynthetic bacterium Rhodopseudomonas palustris WP3–5 using optical fiber-illuminating photobioreactors
.
Biochem Eng J
2006
;
32
:
33
42
. https://doi.org/

[117]

Zarei
Z
,
Malekshahi
P
,
Morowvat
MH
et al.
A review of bioreactor configurations for hydrogen production by cyanobacteria and microalgae
.
Int J Hydrog Energy
2024
;
49
:
472
95
. https://doi.org/

[118]

Giri
DD
,
Dwivedi
H
,
Khalaf
A
et al.
Sustainable production of algae-bacteria granular consortia based biological hydrogen: new insights
.
Bioresour Technol
2022
;
352
:
127036
. https://doi.org/

[119]

Uyar
B
,
Eroglu
I
,
Yucel
M
et al.
Effect of light intensity, wavelength and illumination protocol on hydrogen production in photobioreactors
.
Int J Hydrog Energy
2007
;
32
:
4670
7
. https://doi.org/

[120]

Kim
D-H
,
Kim
M-S.
Hydrogenases for biological hydrogen production
.
Bioresour Technol
2011
;
102
:
8423
31
. https://doi.org/

[121]

Asada
Y
,
Miyake
J.
Photobiological hydrogen production
.
J Biosci Bioeng
1999
;
88
:
1
6
. https://doi.org/

[122]

Nicolet
Y
,
Cavazza
C
,
Fontecilla-Camps
JC.
Fe-only hydrogenases: structure, function and evolution
.
J Inorg Biochem
2002
;
91
:
1
8
. https://doi.org/

[123]

Sivaramakrishnan
R
,
Shanmugam
S
,
Sekar
M
et al.
Insights on biological hydrogen production routes and potential microorganisms for high hydrogen yield
.
Fuel
2021
;
291
:
120136
. https://doi.org/

[124]

Hallenbeck
PC.
Microbial paths to renewable hydrogen production
.
Biofuels
2011
;
2
:
285
302
. https://doi.org/

[125]

Frey
M.
Hydrogenases: hydrogen-activating enzymes
.
ChemBioChem
2002
;
3
:
153
60
. https://doi.org/

[126]

Lee
H-S
,
Vermaas
WF
,
Rittmann
BE.
Biological hydrogen production: prospects and challenges
.
Trends Biotechnol
2010
;
28
:
262
71
. https://doi.org/

[127]

Stiebritz
MT
,
Reiher
M.
Hydrogenases and oxygen
.
Chem Sci
2012
;
3
:
1739
. https://doi.org/

[128]

Ghiasian
M.
Biophotolysis-based hydrogen production by cyanobacteria
.
Prospects Renew Bioprocess Future Energy Syst
2019
;
10
:
161
84
. https://doi.org/

[129]

Markov
SA.
Hydrogen production in bioreactors: current trends
.
Energy Procedia
2012
;
29
:
394
400
. https://doi.org/

[130]

Carolin Christopher
F
,
Kumar
PS
,
Vo
D-VN
et al.
A review on critical assessment of advanced bioreactor options for sustainable hydrogen production
.
Int J Hydrog Energy
2021
;
46
:
7113
36
. https://doi.org/

[131]

Katsuda
T
,
Arimoto
T
,
Igarashi
K
et al.
Light intensity distribution in the externally illuminated cylindrical photo-bioreactor and its application to hydrogen production by Rhodobacter capsulatus
.
Biochem Eng J
2000
;
5
:
157
64
. https://doi.org/

[132]

Ni
M.
An overview of hydrogen storage technologies
.
Energy Explor Exploit
2006
;
24
:
197
209
. https://doi.org/

[133]

Elberry
AM
,
Thakur
J
,
Santasalo-Aarnio
A
et al.
Large-scale compressed hydrogen storage as part of renewable electricity storage systems
.
Int J Hydrog Energy
2021
;
46
:
15671
90
. https://doi.org/

[134]

Zhang
T
,
Uratani
J
,
Huang
Y
et al.
Hydrogen liquefaction and storage: recent progress and perspectives
.
Renew Sustain Energy Rev
2023
;
176
:
113204
. https://doi.org/

[135]

Lim
KL
,
Kazemian
H
,
Yaakob
Z
et al.
Solid‐state materials and methods for hydrogen storage: a critical review
.
Chem Eng Technol
2010
;
33
:
213
26
. https://doi.org/

[136]

Tarasov
BP
,
Fursikov
PV
,
Volodin
AA
et al.
Metal hydride hydrogen storage and compression systems for energy storage technologies
.
Int J Hydrog Energy
2021
;
46
:
13647
57
. https://doi.org/

[137]

Preuster
P
,
Alekseev
A
,
Wasserscheid
P.
Hydrogen storage technologies for future energy systems
.
Ann Rev Chem Biomol Eng
2017
;
8
:
445
71
. https://doi.org/

[138]

He
T
,
Cao
H
,
Chen
P.
Complex hydrides for energy storage, conversion, and utilization
.
Adv Mater
2019
;
31
:
e1902757
. https://doi.org/

[139]

Osman
AI
,
Ayati
A
,
Farrokhi
M
et al.
Innovations in hydrogen storage materials: synthesis, applications, and prospects
.
J Storage Mater
2024
;
95
:
112376
. https://doi.org/

[140]

Bogdanovic
B
,
Spliethoff
B.
Active MgH2 Mg-systems for hydrogen storage
.
Int J Hydrog Energy
1987
;
12
:
863
73
. https://doi.org/

[141]

Liu
W
,
Aguey-Zinsou
K-F.
Low temperature synthesis of LaNi5 nanoparticles for hydrogen storage
.
Int J Hydrog Energy
2016
;
41
:
1679
87
. https://doi.org/

[142]

Dematteis
EM
,
Berti
N
,
Cuevas
F
et al.
Substitutional effects in TiFe for hydrogen storage: a comprehensive review
.
Mater Adv
2021
;
2
:
2524
60
. https://doi.org/

[143]

Zhang
Y
,
Li
C
,
Yuan
Z
et al.
Research progress of TiFe-based hydrogen storage alloys
.
J Iron Steel Res Int
2022
;
29
:
537
51
. https://doi.org/

[144]

Liu
H
,
Zhang
J
,
Sun
P
et al.
An overview of TiFe alloys for hydrogen storage: structure, processes, properties, and applications
.
J Storage Mater
2023
;
68
:
107772
. https://doi.org/

[145]

Li
Y
,
Zhang
Q
,
Ren
L
et al.
Core–shell nanostructured magnesium-based hydrogen storage materials: a critical review. Ind
.
Chem Mater
2023
;
1
:
282
98
. https://doi.org/

[146]

Au
M.
Hydrogen storage properties of magnesium based nanostructured composite materials
.
Mater Sci Eng B
2005
;
117
:
37
44
. https://doi.org/

[147]

Yu
X
,
Tang
Z
,
Sun
D
et al.
Recent advances and remaining challenges of nanostructured materials for hydrogen storage applications
.
Prog Mater Sci
2017
;
88
:
1
48
. https://doi.org/

[148]

Varin
RA
,
Czujko
T.
Overview of processing of nanocrystalline hydrogen storage intermetallics by mechanical alloying/milling
.
Mater Manuf Process
2002
;
17
:
129
56
. https://doi.org/

[149]

Liu
Y
,
Chabane
D
,
Elkedim
O.
Intermetallic compounds synthesized by mechanical alloying for solid-state hydrogen storage: a review
.
Energies
2021
;
14
:
5758
. https://doi.org/

[150]

Zhou
C
,
Zhang
J
,
Bowman
RC
et al.
Roles of Ti-based catalysts on magnesium hydride and its hydrogen storage properties
.
Inorganics
2021
;
9
:
36
. https://doi.org/

[151]

Wronski
ZS
,
Carpenter
G
,
Czujko
T
et al.
A new nanonickel catalyst for hydrogen storage in solid-state magnesium hydrides
.
Int J Hydrog Energy
2011
;
36
:
1159
66
. https://doi.org/

[152]

Rather
S.
Preparation, characterization and hydrogen storage studies of carbon nanotubes and their composites: a review
.
Int J Hydrog Energy
2020
;
45
:
4653
72
. https://doi.org/

[153]

Comanescu
C.
Graphene supports for metal hydride and energy storage applications
.
Crystals
2023
;
13
:
878
. https://doi.org/

[154]

Jain
V
,
Kandasubramanian
B.
Functionalized graphene materials for hydrogen storage
.
J Mater Sci
2020
;
55
:
1865
903
. https://doi.org/

[155]

Zhao
W
,
Fierro
V
,
Zlotea
C
et al.
Optimization of activated carbons for hydrogen storage
.
Int J Hydrog Energy
2011
;
36
:
11746
51
. https://doi.org/

[156]

Mohan
M
,
Sharma
VK
,
Kumar
EA
et al.
Hydrogen storage in carbon materials—a review
.
Energy Storage
2019
;
1
:
e35
. https://doi.org/

[157]

Cabria
I
,
López
MJ
,
Alonso
JA.
Hydrogen storage in pure and Li-doped carbon nanopores: combined effects of concavity and doping
.
J Chem Phys
2008
;
128
:
144704
. https://doi.org/

[158]

Gao
X
,
Zhong
Z
,
Huang
L
et al.
The role of transition metal doping in enhancing hydrogen storage capacity in porous carbon materials
.
Nano Energy
2023
;
118
:
109038
. https://doi.org/

[159]

Wang
S
,
Liu
J
,
Li
Z
et al.
Preparation and properties of hydrogen storage materials of porous carbon composite diatom skeleton
.
Int J Hydrog Energy
2024
;
64
:
773
80
. https://doi.org/

[160]

Elyasi
S
,
Saha
S
,
Hameed
N
et al.
Emerging trends in biomass-derived porous carbon materials for hydrogen storage
.
Int J Hydrog Energy
2024
;
62
:
272
306
. https://doi.org/

[161]

Hu
YH
,
Zhang
L.
Hydrogen storage in metal-organic frameworks
.
Adv Mater
2010
;
22
:
E117
30
. https://doi.org/

[162]

Juan-Juan
J
,
Marco-Lozar
JP
,
Suárez-García
F
et al.
A comparison of hydrogen storage in activated carbons and a metal–organic framework (MOF-5)
.
Carbon
2010
;
48
:
2906
9
. https://doi.org/

[163]

Lin
K-S
,
Adhikari
AK
,
Ku
C-N
et al.
Synthesis and characterization of porous HKUST-1 metal organic frameworks for hydrogen storage
.
Int J Hydrog Energy
2012
;
37
:
13865
71
. https://doi.org/

[164]

Wong
CY
,
Wong
WY
,
Sudarsono
W
et al.
Tuning the functionality of metal–organic frameworks (MOFs) for fuel cells and hydrogen storage applications
.
J Mater Sci
2023
;
58
:
8637
77
. https://doi.org/

[165]

Mulky
L
,
Srivastava
S
,
Lakshmi
T
et al.
An overview of hydrogen storage technologies – key challenges and opportunities
.
Mater Chem Phys
2024
;
325
:
129710
. https://doi.org/

[166]

Li
C
,
Peng
P
,
Zhou
DW
et al.
Research progress in LiBH4 for hydrogen storage: a review
.
Int J Hydrog Energy
2011
;
36
:
14512
26
. https://doi.org/

[167]

Pan
Y.
Exploring the structural, physical properties and hydrogen storage properties of LiBHx(x=1 and 4) lithium borohydrides
.
Ceram Int
2024
;
50
:
3837
42
. https://doi.org/

[168]

Zhu
Y
,
Ouyang
L
,
Zhong
H
et al.
Closing the loop for hydrogen storage: facile regeneration of NaBH4 from its hydrolytic product
.
Angew Chem
2020
;
132
:
8701
7
. https://doi.org/

[169]

Ali
NA
,
Ismail
M.
Modification of NaAlH4 properties using catalysts for solid-state hydrogen storage: a review
.
Int J Hydrog Energy
2021
;
46
:
766
82
. https://doi.org/

[170]

Ding
Z
,
Lu
Y
,
Li
L
et al.
High reversible capacity hydrogen storage through Nano-LiBH4 + Nano-MgH2 system
.
Energy Storage Mater
2019
;
20
:
24
35
. https://doi.org/

[171]

Xiong
R
,
Sang
G
,
Zhang
G
et al.
Evolution of the active species and catalytic mechanism of Ti doped NaAlH4 for hydrogen storage
.
Int J Hydrog Energy
2017
;
42
:
6088
95
. https://doi.org/

[172]

Kaneko
H
,
Miura
T
,
Fuse
A
et al.
Rotary-type solar reactor for solar hydrogen production with two-step water splitting process
.
Energy Fuels
2007
;
21
:
2287
93
. https://doi.org/

[173]

Pellegrini
LA
,
Spatolisano
E
,
Restelli
F
et al.
Toluene/methylcyclohexane as green H2 carrier
. In:
Pernici
B
,
Della Torre
S
,
Colosimo
BM
, et al. (eds.)
,
Green H2 Transport through LH2, NH3 and LOHC. SpringerBriefs in Applied Sciences and Technology
.
Cham
:
Springer
,
2024
,
49
58
. https://doi.org/

[174]

Sisáková
K
,
Podrojková
N
,
Oriňaková
R
et al.
Novel catalysts for dibenzyltoluene as a potential liquid organic hydrogen carrier use—a mini-review
.
Energy Fuels
2021
;
35
:
7608
23
. https://doi.org/

[175]

Liu
H
,
Xue
J
,
Yu
P
et al.
Hydrogenation of N-ethylcarbazole with hydrogen-methane mixtures for hydrogen storage
.
Fuel
2023
;
331
:
125920
. https://doi.org/

[176]

Rao
P
,
Yoon
M.
Potential liquid-organic hydrogen carrier (LOHC) systems: a review on recent progress
.
Energies
2020
;
13
:
6040
. https://doi.org/

[177]

Singh
H
,
Li
C
,
Cheng
P
et al.
A critical review of technologies, costs, and projects for production of carbon-neutral liquid e-fuels from hydrogen and captured CO2
.
Energy Adv.
2022
;
1
:
580
605
. https://doi.org/

[178]

Makaryan
IA
,
Sedov
IV
,
Salgansky
EA
et al.
A comprehensive review on the prospects of using hydrogen–methane blends: challenges and opportunities
.
Energies
2022
;
15
:
2265
. https://doi.org/

[179]

Aziz
M
,
Wijayanta
AT
,
Nandiyanto
ABD.
Ammonia as effective hydrogen storage: a review on production, storage and utilization
.
Energies
2020
;
13
:
3062
. https://doi.org/

[180]

Garcia
G
,
Arriola
E
,
Chen
W-H
et al.
A comprehensive review of hydrogen production from methanol thermochemical conversion for sustainability
.
Energy
2021
;
217
:
119384
. https://doi.org/

[181]

Epelle
EI
,
Obande
W
,
Udourioh
GA
et al.
Perspectives and prospects of underground hydrogen storage and natural hydrogen
.
Sustainable Energy Fuels
2022
;
6
:
3324
43
. https://doi.org/

[182]

Song
L
,
Wang
S
,
Jiao
C
et al.
Thermodynamics study of hydrogen storage materials
.
J Chem Thermodyn
2012
;
46
:
86
93
. https://doi.org/

[183]

Pang
Y
,
Li
Q.
A review on kinetic models and corresponding analysis methods for hydrogen storage materials
.
Int J Hydrog Energy
2016
;
41
:
18072
87
. https://doi.org/

[184]

Zhang
Y
,
Jia
Z
,
Yuan
Z
et al.
Development and application of hydrogen storage
.
J Iron Steel Res Int
2015
;
22
:
757
70
. https://doi.org/

[185]

Li
M
,
Zhu
Y
,
Yang
C
et al.
Enhanced electrochemical hydrogen storage properties of Mg2NiH4 by coating with nano-nickel
.
Int J Hydrog Energy
2015
;
40
:
13949
56
. https://doi.org/

[186]

Xia
G
,
Chen
X
,
Zhao
Y
et al.
High-performance hydrogen storage nanoparticles inside hierarchical porous carbon nanofibers with stable cycling
.
ACS Appl Mater Interfaces
2017
;
9
:
15502
9
. https://doi.org/

[187]

Kukkapalli
VK
,
Kim
S
,
Thomas
SA.
Thermal management techniques in metal hydrides for hydrogen storage applications: a review
.
Energies
2023
;
16
:
3444
. https://doi.org/

[188]

Nguyen
HQ
,
Mourshed
M
,
Paul
B
et al.
An experimental study of employing organic phase change material for thermal management of metal hydride hydrogen storage
.
J Storage Mater
2022
;
55
:
105457
. https://doi.org/

[189]

Tong
L
,
Xiao
J
,
Bénard
P
et al.
Thermal management of metal hydride hydrogen storage reservoir using phase change materials
.
Int J Hydrog Energy
2019
;
44
:
21055
66
. https://doi.org/

[190]

Hassan
IA
,
Ramadan
HS
,
Saleh
MA
et al.
Hydrogen storage technologies for stationary and mobile applications: review, analysis and perspectives
.
Renew Sustain Energy Rev
2021
;
149
:
111311
. https://doi.org/

[191]

Hu
M-L
,
Masoomi
MY
,
Morsali
A.
Template strategies with MOFs
.
Coord Chem Rev
2019
;
387
:
415
35
. https://doi.org/

[192]

Lyu
J
,
Lider
A
,
Kudiiarov
V.
Using ball milling for modification of the hydrogenation/dehydrogenation process in magnesium-based hydrogen storage materials: an overview
.
Metals
2019
;
9
:
768
. https://doi.org/

[193]

Schlapbach
L
,
Züttel
A.
Hydrogen-storage materials for mobile applications
.
Nature
2001
;
414
:
353
8
. https://doi.org/

[194]

Li
X
,
Sun
X
,
Song
Q
et al.
A critical review on integrated system design of solar thermochemical water-splitting cycle for hydrogen production
.
Int J Hydrog Energy
2022
;
47
:
33619
42
. https://doi.org/

[195]

Singh
SK
,
Tiwari
AK.
Solar-powered hydrogen production: advancements, challenges, and the path to net-zero emissions
.
Int J Hydrog Energy
2024
;
84
:
549
79
. https://doi.org/

[196]

Baljit
S
,
Chan
H-Y
,
Sopian
K.
Review of building integrated applications of photovoltaic and solar thermal systems
.
J Clean Prod
2016
;
137
:
677
89
. https://doi.org/

[197]

Abdelsalam
E
,
Almomani
F
,
Azzam
A
et al.
Synergistic energy solutions: solar chimney and nuclear power plant integration for sustainable green hydrogen, electricity, and water production
.
Process Saf Environ Prot
2024
;
186
:
756
72
. https://doi.org/

[198]

Abomazid
AM
,
El-Taweel
NA
,
Farag
HEZ.
Optimal energy management of hydrogen energy facility using integrated battery energy storage and solar photovoltaic systems
.
IEEE Trans Sustain Energy
2022
;
13
:
1457
68
. https://doi.org/

[199]

Mehrjerdi
H.
Off-grid solar powered charging station for electric and hydrogen vehicles including fuel cell and hydrogen storage
.
Int J Hydrog Energy
2019
;
44
:
11574
83
. https://doi.org/

[200]

Egeland-Eriksen
T
,
Hajizadeh
A
,
Sartori
S.
Hydrogen-based systems for integration of renewable energy in power systems: achievements and perspectives
.
Int J Hydrog Energy
2021
;
46
:
31963
83
. https://doi.org/

[201]

Aguirre
M
,
Couto
H
,
Valla
MI.
Analysis and simulation of a hydrogen based electric system to improve power quality in distributed grids
.
Int J Hydrog Energy
2012
;
37
:
14959
65
. https://doi.org/

[202]

Lin
R-H
,
Zhao
Y-Y
,
Wu
B-D.
Toward a hydrogen society: hydrogen and smart grid integration
.
Int J Hydrog Energy
2020
;
45
:
20164
75
. https://doi.org/

[203]

Al-Othman
A
,
Tawalbeh
M
,
Martis
R
et al.
Artificial intelligence and numerical models in hybrid renewable energy systems with fuel cells: advances and prospects
.
Energy Convers Manage
2022
;
253
:
115154
. https://doi.org/

[204]

Sahraie
E
,
Kamwa
I
,
Moeini
A
et al.
Component and system levels limitations in power-hydrogen systems: analytical review
.
Energy Strategy Rev
2024
;
54
:
101476
. https://doi.org/

[205]

Hüner
B.
Mathematical modeling of an integrated photovoltaic-assisted PEM water electrolyzer system for hydrogen production
.
Int J Hydrog Energy
2024
;
79
:
594
608
. https://doi.org/

[206]

Ganjehsarabi
H.
Performance assessment of solar-powered high pressure proton exchange membrane electrolyzer: a case study for Erzincan
.
Int J Hydrog Energy
2019
;
44
:
9701
7
. https://doi.org/

[207]

Schiller
G
,
Lang
M
,
Szabo
P
et al.
Solar heat integrated solid oxide steam electrolysis for highly efficient hydrogen production
.
J Power Sources
2019
;
416
:
72
78
. https://doi.org/

[208]

Souleymane
C
,
Zhao
J
,
Li
W.
Efficient utilization of waste heat from molten carbonate fuel cell in parabolic trough power plant for electricity and hydrogen coproduction
.
Int J Hydrog Energy
2022
;
47
:
81
91
. https://doi.org/

[209]

Ruiming
F.
Multi-objective optimized operation of integrated energy system with hydrogen storage
.
Int J Hydrog Energy
2019
;
44
:
29409
17
. https://doi.org/

[210]

Sikiru
S
,
Oladosu
TL
,
Amosa
TI
et al.
Hydrogen-powered horizons: transformative technologies in clean energy generation, distribution, and storage for sustainable innovation
.
Int J Hydrog Energy
2024
;
56
:
1152
82
. https://doi.org/

[211]

Moreno Escobar
JJ
,
Morales Matamoros
O
,
Tejeida Padilla
R
et al.
A comprehensive review on smart grids: challenges and opportunities
.
Sensors
2021
;
21
:
6978
. https://doi.org/

[212]

Bakkar
M
,
Bogarra
S
,
Córcoles
F
et al.
Artificial intelligence-based protection for smart grids
.
Energies
2022
;
15
:
4933
. https://doi.org/

[213]

Sai Ramesh
A
,
Vigneshwar
S
,
Vickram
S
et al.
Artificial intelligence driven hydrogen and battery technologies – a review
.
Fuel
2023
;
337
:
126862
. https://doi.org/

[214]

Hayati
MM
,
Safari
A
,
Nazari-Heris
M
et al.
Hydrogen-Incorporated Sector-Coupled Smart Grids: A Systematic Review and Future Concepts
. In:
Vahidinasab
V
,
Mohammadi-Ivatloo
B
,
Shiun Lim
J
(eds.)
,
Green Hydrogen in Power Systems
.
Green Energy and Technology. Cham
:
Springer
,
2024
,
22
58
. https://doi.org/

[215]

Koepf
E
,
Zoller
S
,
Luque
S
et al.
Liquid fuels from concentrated sunlight: an overview on development and integration of a 50 kW solar thermochemical reactor and high concentration solar field for the SUN-to-LIQUID project
AIP Conf. Proc.
2019
;
2126
:
180012
. https://doi.org/

[216]

Schäppi
R
,
Rutz
D
,
Dähler
F
et al.
Drop-in fuels from sunlight and air
.
Nature
2022
;
601
:
63
68
. https://doi.org/

[217]

Djizanne
H
,
Murillo Rueda
C
,
Brouard
B
et al.
Blowout prediction on a salt cavern selected for a hydrogen storage pilot
.
Energies
2022
;
15
:
7755
. https://doi.org/

[218]

Lyakhovskaya
VR
,
Magomadov
IA
,
Balhasan
SA
et al.
Hydrogen storage opportunities in UAE: potential innovations and advances
.
Meeting Abstracts
2024
;
MA2024-11
:
4
7
,
Abu Dhabi,UAE
ADIPEC, 2024
. https://doi.org/

[219]

Stetson
NT
,
Allendorf
MD
,
Adams
J
et al.
(Invited) an overview of the hydrogen materials - advanced research consortium (HyMARC), a DOE energy materials network consortium, to accelerate development of hydrogen storage materials
.
Meet Abstr
2017
;
MA2017-01
:
1663
1663
. https://doi.org/

[220]

Hara
D.
(Invited) Japan’s current policy and activity toward hydrogen society-introduction of representative
projects in Japan
. In: Daishu Hara 2020, Meeting Abstracts, MA2020-02, p.
2502
.
Pennington, NJ, USA
:
The Electrochemical Society
,
2020
. https://doi.org/

[221]

Rashidi
S
,
Karimi
N
,
Sunden
B
et al.
Progress and challenges on the thermal management of electrochemical energy conversion and storage technologies: fuel cells, electrolysers, and supercapacitors
.
Prog Energy Combust Sci
2022
;
88
:
100966
. https://doi.org/

[222]

Murphy
CA
,
Schleifer
A
,
Eurek
K.
A taxonomy of systems that combine utility-scale renewable energy and energy storage technologies
.
Renew Sustain Energy Rev
2021
;
139
:
110711
. https://doi.org/

[223]

Staffell
I
,
Scamman
D
,
Velazquez Abad
A
et al.
The role of hydrogen and fuel cells in the global energy system
.
Energy Environ Sci
2019
;
12
:
463
91
. https://doi.org/

[224]

Ehteshami
SMM
,
Chan
SH.
The role of hydrogen and fuel cells to store renewable energy in the future energy network – potentials and challenges
.
Energy Policy
2014
;
73
:
103
9
. https://doi.org/

[225]

Juma
SA
,
Ayeng’o
SP
,
Kimambo
CZM.
A review of control strategies for optimized microgrid operations
.
IET Renew Power Gen
2024
;
18
:
2785
818
. https://doi.org/

[226]

Petrollese
M
,
Valverde
L
,
Cocco
D
et al.
Real-time integration of optimal generation scheduling with MPC for the energy management of a renewable hydrogen-based microgrid
.
Appl Energy
2016
;
166
:
96
106
. https://doi.org/

[227]

Hossain
MB
,
Islam
MR
,
Muttaqi
KM
et al.
A compensation strategy for mitigating intermittencies within a PV powered microgrid using a hybrid multilevel energy storage system
.
IEEE Trans Ind Appl
2023
;
59
:
1
12
. https://doi.org/

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.