Next Article in Journal
Current Scenario of Solar Energy Applications in Bangladesh: Techno-Economic Perspective, Policy Implementation, and Possibility of the Integration of Artificial Intelligence
Next Article in Special Issue
Hybrid PEM Fuel Cell Power Plants Fuelled by Hydrogen for Improving Sustainability in Shipping: State of the Art and Review on Active Projects
Previous Article in Journal
Optimal Control of a Single-Stage Modular PV-Grid-Driven System Using a Gradient Optimization Algorithm
Previous Article in Special Issue
Functional Model of Power Grid Stabilization in the Green Hydrogen Supply Chain System—Conceptual Assumptions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Global Review of the Hydrogen Energy Eco-System

by
Omer Faruk Noyan
1,
Muhammad Mahmudul Hasan
2 and
Nezih Pala
2,*
1
The Center for Ethnic and Cultural Studies, Fort Myers, FL 33967, USA
2
Electrical & Computer Engineering, Florida International University, Miami, FL 33174, USA
*
Author to whom correspondence should be addressed.
Submission received: 2 December 2022 / Revised: 3 January 2023 / Accepted: 30 January 2023 / Published: 2 February 2023
(This article belongs to the Special Issue Advances in Hydrogen Energy Ⅱ)

Abstract

:
Climate change primarily caused by the greenhouse gases emitted as a result of the consumption of carbon-based fossil fuels is considered one of the biggest challenges that humanity has ever faced. Moreover, the Ukrainian crisis in 2022 has complicated the global energy and food status quo more than ever. The permanency of this multifaceted fragility implies the need for increased efforts to have energy independence and requires long-term solutions without fossil fuels through the use of clean, zero-carbon renewables energies. Hydrogen technologies have a strong potential to emerge as an energy eco-system in its production-storage-distribution-utilization stages, with its synergistic integration with solar-wind-hydraulic-nuclear and other zero-carbon, clean renewable energy resources, and with the existing energy infrastructure. In this paper, we provide a global review of hydrogen energy need, related policies, practices, and state of the art for hydrogen production, transportation, storage, and utilization.

1. Introduction

Anthropogenic CO2 emissions that cause global warming have been increasing at a worldwide scale since the end of 19th century. With a combined 66.4% of worldwide fossil fuel use and 67.8% of global fossil CO2 emissions in 2021, China, the United States, the European Union, Japan, Russia, and India continued to lead the world in CO2 emissions. India and Russia had the biggest relative increases in fossil CO2 emissions among the six countries in 2021 in contrast with 2020 [1]. This situation has motivated the search for clean energy solutions. As one of the most important alternatives, hydrogen (H2) has a long history of being used as fuel, dating back to 1806 when the first internal combustion engine (ICE) was powered by a mix of hydrogen and oxygen. With the increasing global environmental concerns, hydrogen has recently emerged as a major secondary energy source to store and transport energy produced from other sources, since it can be used without direct emissions of air pollutants or greenhouse gases. Hydrogen, as a synthetic fuel, can be produced from a variety of low-carbon sources, such as wind and solar, and can store energy efficiently with high energy content per unit mass, which further contributes to reducing the environmental impact. Developing hydrogen technologies are expected to provide a pathway to a resilient and sustainable energy eco-system by: (1) using hydrogen as an efficient energy storage and transport medium to replace carbon-based fuels; (2) storing the electricity to meet weekly or monthly imbalances in supply and demand in renewable energy (RE) systems; and (3) replacing the carbon-intensive hydrogen production methods with green ones to supply for the demand of all industries [2].
Policies have been developed at national and global levels to incentivize and support research, development and adoption to achieve the potential of hydrogen technologies. The recent geopolitical conflicts have deeply impacted global energy security, and are expected to lead to the revision of such policies. In this paper, we provide a global review of hydrogen energy need, related policies, practices, and state of the art for hydrogen production, transportation, storage, and utilization.

2. Overview of the Global Energy Landscape

After the world economy had been affected by COVID-19 pandemics, the impacts of the Russia–Ukraine conflict in 2022 on international energy supplies are causing further vital consequences. Russia is one of the largest fossil fuel producers in the world, being third in the production and exporting of crude oil; second in the production and the largest in the export of natural gas (NG); and the third in the export of coal. It is the producer of 12% of the world’s oil demand, 17% of the NG demand, and 5% of the coal demand and deeply integrated into Europe’s distribution networks. The European Union (EU) and United Kingdom, combined, import about 40% of the NG, about 35% of the crude oil and more than 45% of the coal from Russia [3,4]. The Russia–Ukraine conflict and sanctions have increased the interest in H2 as a potential cost-efficient, clean, renewable, sustainable fuel for the future.
In addition to its impact on fossil fuels, the Russia–Ukraine conflict also affected the biofuel ecosystem. Russia is the world’s largest exporter of wheat and the second largest exporter of sunflower oil, while Ukraine is the world’s largest exporter of sunflower oil, the fourth largest exporter of maize, and the fifth largest exporter of wheat. Russia and Ukraine produce nearly 25% of the world’s wheat and barley, 20% of the world’s corn, and over 60% of its sunflower oil [5,6]. Rising food prices due to the conflict raised the concerns for potential famine risk and rekindled the “food versus biofuel” argument. In the US, the leading biofuels producer, 36% of total corn production was used as biofuels (blended into petroleum) in 2021. In the EU, 12 million tons of grain, including wheat and maize, is turned into ethanol—around 7% of the bloc’s production [6]. It was estimated that the calories diverted to biofuel production under the current policies and future commitments are equivalent to the annual needs of 1.9 billion people [7].
The EU has developed a strategy concentrating on emission-free renewable H2 and plans to install 40 gigawatts of green H2 electrolyzer capacity by 2030. Germany recently started a collaboration effort with Australia to decrease the production cost of renewable H2 and accelerate the innovation process in both countries and to minimize reliance on Russian natural gas [8]. China identified H2 as one of the six industries of the future in its 5-year economic plan [9,10]. Japan aims at having 800,000 H2 fuel cell vehicles (FCVs) and 900 H2 refueling stations by 2030 [11]. South Korea has a target to supply 10% of its energy demands with H2 by 2030, increase this rate to 30% before 2040 and make it the country’s largest energy carrier by midcentury [12]. India sees H2 as the opportunity for “quantum leap” towards energy independence by 2047 and shapes its energy policy accordingly [12]. The US Department of Energy (DoE) plans to invest $9.5B for the commercialization of innovative green H2 production technologies with the goal of reducing the production cost by 80% within the next decade. Australia, Saudi Arabia, Morocco, United Arab Emirates and Oman have announced strategies to develop clean H2 to diversify their energy portfolios [13].
Important developments in H2 production, storage, distribution, and utilization have been demonstrated in recent years. As a fuel, H2 has considerably higher energy per unit weight than other fuels such as gas, diesel, and methanol (see Figure 1) [14]. Therefore, it has an immense market potential for the transportation sector, and industry. H2 storage is one of the key factors in the H2 economy. RE electricity is widely stored in chemical batteries but the limited capacity, lifetime, and high cost of batteries have limited their applications. H2 could be employed as an energy carrier, like a battery, for RE electricity. The stored H2 can be used for electrical power production in a gas turbine or fuel cell, or mechanical power production in an ICE. Solid-state materials in standard temperature and pressure conditions (STP) are seen as a promising method of storing H2, especially for mobile systems. In terms of H2 utilization, the deployment of fuel cell (FC) systems opened a new window to H2 fuel success in the transportation sector. Currently, H2-powered engines based on proton exchange membrane (PEM) FCs are deployed worldwide.
H2 is defined as a clean energy source but is mostly produced via the combustion of fossil fuels (grey or black-brown category). For commercial production of H2, four methods come forward, three of which use fossil fuels: steam methane reformation, oxidation, and gasification. The fourth one is electrolysis [15]. When the electricity for H2 production is supplied by clean electricity, (solar photovoltaics -PV-, concentrating solar power, wind, wave, hydropower, and geothermal), the produced H2 is regarded as zero-carbon green H2. This means cleanliness, inexhaustibility, and independence from geopolitical effects. Today, there are highly efficient electrolyzers to produce H2 as an environmentally friendly replacement for fossil fuels in all sectors. The integrated RE H2 system and economy is a supply chain from production to storage, transportation/distribution and utilization (Figure 2). In an established large-scale H2 infrastructure the most featured difficulties are cost reduction and the scaling up of this chain.

3. Renewable Energies

RE globally and technically promises more zero-carbon action. Hydropower and three intermittent energies, solar, wind, and tidal are resources that can forever be used to produce and store H2 via the electrolysis of water, which is also a medium of energy storage. The global RE generation capacity has reached 2.8 TW as of 2021. Within this capacity, hydropower accounts for 1.3 TW while solar and wind powers account for 126 GW and 110 GW, respectively. The share of RE in the global electricity generation increased to 29% in 2020, up from 27% in 2019 [15]. The costs for RE continued to fall in 2021 with a 15% and 13% reduction in the cost of electricity from onshore wind and offshore wind, respectively, and a 13% reduction in the cost of solar PV generation compared to 2020, despite the rising materials and equipment costs. In 2021, about 60% of newly installed RE power had lower costs than the world’s cheapest coal-fired option in the G20 countries. The benefit from RE in 2022 has been unparalleled, given the fossil-fuel price crisis: (i) in Europe, new solar and wind power added in 2021 had lifetime costs per kWh that, on average, were four to six times lower than the 2022 minimal producing expenses of fossil fuels; (ii) the additional RE capacity developed in 2021 may globally lower the cost of producing electric power in 2022 with a minimum of 55 billion USD; and (iii) solar and wind power alone prevented the import of fossil fuels, mostly NG, in the amount of 50 billion USD in Europe during January to May 2022 [16]. As a historical turning point, the 2022 crisis showed that without RE, conditions, particularly in the developed countries, would be much more challenging for consumers, economies, and the environment. Utility-scale solar PV and wind turbines have become effectively competitive over fuel and CO2 cost in Europe.
Solar and wind energy, with their short project lead times and relatively lower initial investment, present affordable pathways for countries to quickly reduce and eventually phase out fossil fuels while limiting the economic impact. The recent report “Renewable Power Generation Costs in 2021” published by the International Renewable Energy Agency (IRENA) states that “RE is by far the cheapest form of power today. 2022 is a stark example of just how economically viable new RE power generation has become. It frees economies from volatile fossil-fuel prices and imports, curbs energy costs and enhances market resilience. Today’s situation is a devastating reminder that RE and energy saving are the future” [16].

4. Clean RE for an Integrated H2 Energy System

H2 is the most abundant and lightest element (2.016 × 10−3 kg/mol) in the universe, and energy-efficient fuel on the Earth (calorific or heating value). It is the richest in energy per-unit mass (140 MJ/kg), and almost three times higher than solid fuels (50 MJ/kg) (see Figure 1). It exists mainly in compounds such as water, methane, H2S, living organisms, organic waste, etc. It is environmentally benign, since water is the only emission product at conversion to energy, and unlike electricity, it can be stored. RE H2 produced either at a central production plant or onsite at a H2 refueling station (HRS) could help achieve carbon-free energy economy. Despite the fact that onsite H2 production by electrolysis typically has higher costs than production at central plants due to limited production capacity, it was shown that the at-the-pump prices end up being similar due to the additional H2 transport cost for the case of central production. H2 production cost constitutes the largest portion of the levelized H2 price at the pump due to the high capital investment in both RE electricity generation and water electrolysis. As the relevant technologies advance, the at-the-pump H2 price is estimated to fall substantially. In a 2022 study, the future at-the-pump prices of RE H2 produced onsite at an HRS in the 26 EU countries are estimated as 11 euros/kg in 2020, decreasing to 5 euros in 2050 while the European Commission (EC) target is 1.8 euros/kg H2 by 2030. With a 3 euro/kg H2 subsidy for RE H2 production, it is expected that the price targets can be reached earlier. Moreover, the subsidy can enable RE H2 to reach an energy-based cost parity with diesel fuel (0.034 euros/MJ for H2 compared to 0.038 euros/MJ for diesel) by 2030. Furthermore, since H2 FCVs being more efficient, they can travel 1.3 times further compared to diesel trucks using the same amount of energy. Therefore, considering both increasing diesel prices and vehicle efficiency, the 3 euros/kg subsidy can pave the way to make H2 cost-competitive against diesel well before 2030. Although subsidies could be a promising way to lower prices, to achieve true climate benefits, robust regulations on sourcing electric power are necessary in addition to direct financial support for RE H2 production [17,18,19,20,21,22].
It has been shown that integrated PV-battery- H2 systems can sustain an affordable electricity cost over the system’s lifetime owing to the energy storage components. Similarly, a multi-optimization problem has been studied for integrated wind-PV energy systems for RE H2 production to assess its exergo-economic performance and to determine the optimal operation conditions [23]. However, when the H2 is integrated with other RE systems the overall complexity of the combined system increases the uncertainties in the power load which requires careful design to address the nonlinear characteristics and added variables [24,25].

5. H2 Production

Though hydrogen is the most common element constituting ~70% of the universe, free hydrogen is not readily available on earth. It is typically bound into molecules such as water and hydrocarbons, particularly methane (CH4). It can be separated from these molecules by using energy in various forms. H2 is identified with a spectrum of colors depending on its source or production method: (1) black-brown (from coal), (2) gray (hydrocarbons without carbon capture and storage—CCS), (3) blue (hydrocarbons with CCS), (4) turquoise (pyrolysis), (5) pink (nuclear powered electricity), (6) yellow (solar energy), (7) red (high-temperature catalytic process), (8) white (naturally-occurring geological hydrogen) and (9) green (clean RE electricity) (Figure 3). Natural gas (NG), crude oil, coal, and water electrolysis processes are the most commonly used sources for H2 production, with shares of 49, 29, 18, and 4%, respectively [26].

5.1. Black-Brown and Gray H2 Production

By far the most widely used H2 production method is the steam reforming of natural gas, also known as steam methane reformation (SMR). This is one of the cheapest H2 production technologies in terms of operation and production costs [27], and is also renowned for its high efficiency operation [28]. Steam and natural gas are heated together up to ~900 °C at high pressure in the presence of a nickel-based catalyst. The process yields a mixture of carbon monoxide (CO) and H2 known as syngas which is further treated in the water–gas shift reaction to produce more hydrogen and carbon dioxide. The H2 produced by this method is named gray H2. However, it has serious downsides, for instance: (1) as an endothermic reaction it requires huge amount of heat supply at extremely high temperature; (2) generating heat to produce extra steam (exceeds the stoichiometric requirement by two to three times) which is required for the curving of deactivation events like coking and moving the equilibrium position of the reaction more in the direction of the H2 generation, and (3) complex extraction of H2 from a combination of other outcomes (CO, CO2, CH4, and H2O) by employing pressure swing adsorption (PSA, separates gases in a mixture of gases to obtain a high degree of purity) and water–gas shift reactors [29]. In a similar process called gasification, steam and oxygen are used to break molecular bonds in black coal or lignite (brown coal) to produce syngas consisting of CO, H2, CO2, CH4, and water vapor. These processes also result in substantial CO2, which must be captured and stored (CCS). To address these limitations, a carbon-efficient process concept was developed that converts captured CO2 to CO2 in two steps. It prevents direct contact between O2 and CH4, lowering safety risks. The external energy demand is reduced by coupling endothermic and exothermic reactions, and the operating temperature to produce H2 is reduced by more than 150 K when compared to conventional CH4 reforming. Iron, nickel, and calcium oxides are used in this process as solid intermediates, which operates between 923 and 998 K. By simultaneously delivering steam and methane to the materials in a fixed bed, an H2-rich stream is created in the first phase, and a CO-rich stream is created in the second step by renewing the materials with a combination that is like air. H2 concentrations of over 65 mol% and CO2 conversion rates of over 80% were reached in proof-of-concept tests. According to cyclical experiments, carbon deposition is reduced, and the ratio of generated H2 to CO is around 2. Experimental findings demonstrate that the process idea enables conventional methane reforming for the generation of syngas to avoid the limitations of thermodynamic equilibrium [29].
Biomass is another source of feedstock for H2 production. As gasification is a well-established and commercially available technology, it is commonly applied to biomass like coal. The oxidation eventually results in a combination of CO, H2, CO2, CH4, N2, and higher hydrocarbons gases [30]. After the creation of syngas H2/CO2 a WGS reactor is used for H2 production with high temperature [30]. It is really hard to obtain pure H2 as biomass produces large amounts of tar even with high temperature operation due to higher hydrocarbons. The efficiency of the biomass ranges from 35–50% [31]. The moisture contents in biomass limit the efficiency of this process as it needs to be vaporized. Gasification of biomass for H2 production are not yet commercially established due to high operation costs and the complicated process for cleaning the H2 from tar.
Approximately 6% of the globally extracted NG and 2% of the globally extracted coal are used for H2 production. Over 95% of current H2 production (black-brown and grey) leading over 2% of global annual CO2 emissions by 2020 is by steam reforming of NG, its partial oxidation, and coal gasification; very little of it is green. Producing H2 in pure form for industry is energy intensive. The global demand for H2 is for its utilization in chemical production (40%), oil refineries (33%), metallurgical industries (3%), and the rest is used in synthetic fuel and petrochemical production, semiconductor and glass manufacturing, welding processes, powering FCVs, and in the food industries [32,33].

5.2. Blue H2 Production

H2 produced from NG processing coupled with CCS is referred to as blue H2 being carbon neutral. Blue H2 production is predicted to reach 80 million metric tons by 2050 [34]. The climate change impact of H2 produced from fossil fuels coupled with CCS is still debated, especially in regard to the leakage risks of indefinite long-term storage of methane gas [35]. H2 produced using fossil-fuel feedstock causes greenhouse gas (GHG) emissions, even when CCS is used. These emissions could be substantial, and the cost of CCS is higher than frequently assumed. The percentages of GHG releases and those of CCS during blue H2 production process determine the impacts of blue H2 on the climate and the competitiveness of blue H2 regarding global environment seems to depend on high CCS and negligible GHG releases. However, according to a research analysis the positive effects of blue H2 production in terms of GHG emissions and climate change are insignificant and the use of blue H2 seems challenging to justify on climate grounds [36]. Establishing H2 supply chains on the basis of fossil fuels, as many national strategies forecast, may be incompatible with decarbonization goals [18]. Therefore, developing advanced supply chains in combination with high CCS capture rates is vital for blue H2 to be a sustainable option for an actual net zero hydrogen economy [37].
In a study of optimizing blue hydrogen production for future energy systems, a northern European perspective based on Germany was investigated [38]. Blue routes provide significant benefits in the hydrogen industry but limited gains in the electricity sector. Systems using CCS (CO2 Capture and Storage) become significantly less expensive than those depending just on renewables in an energy system of the future when vast amounts of carbon-free fuels are required for usage in sectors like long-distance transportation and industrial. Another conclusion is that paths that allow CCS would prioritize carbon-free fuels above electrification. The study found that when conventional CCS technology were used, system costs decreased by 29% and CO2 emissions decreased by 106% (biomass’s negative emission), with hydrogen generated for less than the cost of power. With the possibility for integrated power cycles that enable variable power and hydrogen generation, advanced blue hydrogen technologies might provide an extra 12% decrease in system costs with extra negative emissions because of extremely high CO2 capture metrics. Since handling the intermittent hydrogen fluxes resulting from such an operational strategy proved to be expensive, scenarios with lower variable renewable energy shares proved to be more cost-effective. Originally, these concepts were created to better integrate higher shares of wind and solar energy [38].

5.3. Turquoise H2 Production

Like gray and blue H2, turquoise H2 also uses methane as a feedstock, but the process is driven by heat produced with electricity rather than through the combustion of fossil fuels. In the process called pyrolysis, hydrogen is separated from methane (i.e., NG) in one step via flow through a molten metal catalyst in a bubble column at higher temperatures (1065 °C). This pyrolysis process is an endothermic reaction where two molecules of H2 in gaseous form and one molecule of carbon in solid form is produced from one molecule of methane. This reaction is advantageous over methane reforming as it can be performed in near ambient pressure. Theoretically, this reaction produces a larger amount of solid carbon than H2 with the ratio of 1 to 3 in favor of carbon. However, there is no GHG-free methane reform process. Industrial quality solid carbon can be used as manufacturing feedstock, for example for carbon-fiber production or landfilled which are simpler than capturing and sequestering CO2. Since it is not released into the atmosphere and does not pollute groundwater in landfills, turquoise H2 may contribute to low-carbon H2 production in the future. Pyrolysis has not been fully commercialized yet for H2 production. There are variants of the process that are being developed including thermal, plasma, and catalytic decomposition. The required process temperature for thermal, plasma, and catalytic decomposition are over 1000 °C, over 2000 °C, and below 1000 °C. Plasma processes realized on industrial scale are still being further developed. The other processes are at an early stage of development [39].

5.4. Pink H2 Production

In a global zero-carbon energy system, nuclear-produced H2 has a potential place. In this alternative method, H2 can be produced by steam of which heat source is nuclear power plant. Nuclear energy can be used, as a source of electricity and heat, to produce H2 efficiently and with little to no CO2 emissions.
The nuclear alternative has had the challenge of high cost in the past, but that is now changing in the midst of a global energy crisis causing significant price hikes for fossil fuels as well as making global supplies less secure [40]. Nuclear power is the second largest low-carbon source of electricity after hydropower, supplying 10% of the world’s electricity. However, nuclear power generating capacity will need to more than double to achieve net zero by 2050 [41].
Light water reactor technology presents great potential for H2 production. The reactors reach an operating temperature of ~300 °C, while district heating and seawater desalination processes require ~150 °C. Numerous district heating systems have been in operation for decades. This heat could also be utilized for the production of fresh water, H2 or other products in cogeneration (simultaneous production of electricity, heat and heat-derivative product) [41,42]. Electricity and heat produced at temperature levels of 300 °C in small modular reactors can be used in solid-oxide electrolyzer cells (SOEC). Six small modular reactors with a total capacity of 300 MW can meet the annual H2 demand of a mid-sized ammonia plant (73,000 tons of H2/year) [43]. In the longer term, advanced nuclear reactors will be able to provide the higher temperatures required for industrial processes such as steel and cement production, as well as H2 production by thermochemical water splitting (with some reactor designs having coolant outlet temperatures of 800–1000 °C). A single 1000 MW nuclear power reactor could produce more than 200,000 tons of H2 each year to fuel more than 400,000 FCVs or more than 16,000 long haul FC trucks [44]. In the near future, small-to medium-sized reactors can be dedicated, on an industrial scale, to making H2, which also takes away the cost of storage and transportation. Moreover, using base-load surplus of nuclear or any RE power grid H2 can be produced for the local market by a mix of electrolysis and thermal processes. Nuclear power, by its potential for further development, can be a part of efforts to achieve global climate targets, and it is a sufficiently experimented alternative which can stimulate clean H2 production preferences.

5.5. Green H2 Production

H2 produced by water electrolysis using electricity from clean RE, namely, the green H2, promises to help meet global energy demand while contributing to climate action goals, and represents the most promising energy carrier for the zero-carbon economy. Falling RE energy prices coupled with the declining cost of electrolyzers and increased efficiency due to technology advancements have increased the commercial feasibility of green H2 production. Assuming these costs continue to fall, green H2 can be produced for $0.70–$1.60/kg by 2050, a price range competitive with NG [45].
In 2003, the European Commission (EC) suggested that the EU should attain a H2-based economy in 2050 and estimated that 35% of newly produced vehicles will be fueled by zero-carbon H2 in 2040 [46]. This strategy updated in 2022 aims to ensure low-carbon H2 technologies be able to be deployed at large scale to cover all hard-to-decarbonize sectors. Projects varying from 20 MW to above 100 MW are being developed with the current H2 costs of 5–8 €/kg. The EC launched a call for a 100 MW electrolyzer to scale up manufacturing facilities to multi-GW per annum, to attain the goal of 1.5 $/kg of green H2 by 2025 provided that low-cost RE power is ensured. In this way, it is believed that green H2 production cost parity (or even superiority) with fossil- fuels could be achieved as early as 2025. Deploying zero-emission H2 vehicles (heavy-duty FC H2 buses, trucks, and trains) is an essential part of EU’s H2 policy. They will require a very reliable, high-capacity HRS capable of delivering several tons each day. The regulation requires one HRS available every 150 km along the Trans-European Transport Network and in every urban node by 2030. In order to fully unleash the potential of H2 technologies and establish them as a mainstream means of decarbonization in all transportation modes, cost reduction, increased performance and lifetime, sustainability, recycling, and eco-design in transport FC system components and vehicles are needed [47]. Another H2 -hub plan approved by the EC in July 2022 allocates €5.4 billion to support 41 technology development projects. A small hub near Hamburg, Germany, is already under construction, and a larger hub is being built at the Port of Rotterdam in Netherlands. Portugal will produce green H2 using wind power and transport it into the Port of Rotterdam. Spain houses Europe’s largest green H2 facility with a 100 MW solar farm. In USA, the Intermountain Power Project will also have a 220 MW electrolysis system based on PV generators to produce H2 on site, along with the facilities to store up to 300 GW hours of the gas in underground salt domes [48]. Studies on China’s potential of green H2 production showed that the efficiency of the H2 production from wind power is significantly higher than that from solar power. These efficiencies and green H2 production in each province of China are expected to significantly increase by 2030 [49]. In a recent study in Sweden the feasibility of producing green H2 using electrolysis with RE electricity in HRS was investigated. With hourly solar radiation and wind speed data, and electricity price, simulations were used to assess the cost components of H2 production. The study revealed that integrating the electricity grid in green H2 production was important, wind speed was critical for cost reduction, whereas solar radiation had less influence. Further, a combination of solar and wind could provide better performance in an off-grid scenario. The most encouraging finding was the cost, which was competitive with reported costs in other EU countries [50].
In addition to the currently existing technologies, efficient and direct conversion of sunlight through a photocatalytic process to split ocean water without use of electricity can pave the way to a clean production route. A unique, CdS/ZnTHPP binary nanocomposite was recently created and employed as a reliable photocatalyst for the generation of H2 in the presence of artificial sunlight which is all solid state. It demonstrated a conversion rate of almost ten times greater than that of pure CdS nanorods, the data showed that this nanosystem had remarkable photocatalytic activity. This innovation may lead to further improvement toward high-efficient photocatalytic green H2 productions for real applications [51].
The inherent production intermittency of RE sources makes operation of energy systems with large RE components reliably and securely quite challenging. In an energy system with 100% RE the infrastructures for carbon-free H2 production by electrolysis, and storage and/or transportation can overcome the intermittency of RE. The commercial production of H2 by electrolysis of water can achieve efficiency up to 81%. Less than 0.1% of H2 production globally comes from water electrolysis today, and the H2 produced by this means is mostly used in markets where high-purity H2 is necessary (electronics and polysilicon), but electrolysis, as an energy-intensive process for H2 production, is still confronting challenges [52,53,54,55,56,57]. Electrolysis can be classified into three types: (i) alkaline (AWE), (ii) proton-exchange membrane (PEM), and iii) solid-oxide steam (SOSE). H2 production by AWE is well established technology up to the megawatt range for commercial level (Figure 4a).
PEM electrolyzers have the ability to deliver on-site and on-demand H2 at very high purity and carbon-free (Figure 4b). Solid polymer electrolyte membrane used in electrolyzer is the same as proton exchange membrane. They use pure water, and therefore do not need to deal with the recovery and recycling of the potassium hydroxide electrolyte solution that is necessary with AWE. They are also relatively small and more appealing than AWE in dense urban areas. They can produce and store highly compressed H2 at HRS and provide frequency reserves.
SOSE process is a promising technology with high scalability, high application flexibility, pure H2 production, enhanced reaction kinetics at higher temperatures with low capital costs, and high degree of electrical efficiency [59]. It requires a heat source because steam is the electrolysis medium. A SOSE’s heat source options include nuclear power plants, solar thermal systems, and geothermal energy. In contrast to AWE and PEM, an SOSE has the ability to function in reverse mode as an FC, turning H2 back to electric power. As a result, it might work in conjunction with H2 storage facilities to offer stabilizing benefits to the grid. Managing the material degradation brought on by the extreme operation temperatures is one of the major drawbacks that the SOSE electrolyzers are facing [2]. As a way to generate oxygen for both human nourishment and liquid oxygen rocket fuel, MIT demonstrated an SOSE system in the Mars Oxygen ISRU Experiment on the Perseverance rover in 2014. In April 2021, NASA claimed it has produced, 4–5 g of oxygen on Mars from CO2 in the Mars atmosphere using SOSE [60].
Today’s total global H2 production of 70 Mt would require 617 million m3 water if it all is produced by electrolysis. It corresponds to 1.3% of the water consumption of the global energy sector [2]. However, in some regions, there may be discordance between localities of suitable water resources and those of RE sources, and this may pose a serious obstacle to H2 production. To address this challenge, H2 production from the air can offer a solution. Freshwater may be collected from the environment employing a hygroscopic electrolyte driven by solar or wind with a current density up to 574 mA cm−2 and a consistent quality at a Faradaic efficiency of 95%. This direct air electrolysis (DAE) can work in arid environments with a relative humidity (RH) of 4%, overcoming water supply challenges and producing green H2 sustainably. There are 12.9 trillion tons of water in air at any moment. Even in the Sahel desert, the average RH is about 20%, and at Uluru desert (Australia) is 21%. DAE systems may be expanded to readily create H2 in far-flung, dry, and dispersed regions [61]. In coastal locations, seawater could replace other sources. Reverse osmosis for desalination has little effect on the overall expenditures of water electrolysis, raising total H2 generation costs by $0.01–0.02/kg H2 with an electricity requirement of 3–4 kWh/m3 of water [62]. However, seawater causes corrosion and chlorine byproducts. Future research can make this method more feasible and at a lower cost. In electrolysis, the solar-powered system can be improved by applying magnetic, ultrasonic, pulsating electric or light energy fields [63]. It is known that these applications enhance the efficiency of electrolysis. Further studies are needed to use the physicochemical and kinetic properties of water for this innovative approach to be developed.
Recently, semiconductor-based photocatalysis and photoelectrochemical water splitting have emerged as novel and promising techniques for solar hydrogen production. In these methods, a semiconductor material is exposed to light to generate electron–hole pairs which undergo redox reactions resulting hydrogen production. The basic principle of these two processes is the same. As the first step, the semiconductor material absorbs photons from the sunlight. In the second step, an electron-hole pair is generated and is separated. After that, the created positive and negative charges are transported with the help of the catalysts. Finally, surface chemical reactions take place [64]. However, there are some dissimilarities between photocatalysis and photoelectrochemical processes. The semiconductor material in photocatalysis is in a powder form whereas it is in a thin film form for photoelectrochemical processes. The charge transfer path is relatively shorter in photocatalytic system them in photoelectrochemical system. Considering everything, the photoelectrochemical system has the upper hand over the photocatalytic system in terms of cost and efficiency as well [65]. One of the major challenges in photocatalysis is the recombination of photo-generated charge carriers before they reach the surface for the reduction and oxidation reactions, which reduces the quantum efficiency [66]. It was shown that introduction of heterojunction materials could substantially improve conversion efficiency. For example, Ag/WO3/Bi2WO6 heterojunction yielded improved conversion rate by up to 250% and 190% compared to pristine Bi2WO6 and WO3/Bi2WO6 samples, respectively. The improvement was attributed to the increased separation efficiency of the photogenerated electrons and holes as well as the improved visible-light absorption, both provided by the heterojunction [67]. Studies have shown that using noble materials such as Ru, Rh, and Ta in the photoactive semiconductor provides excellent photocatalytic performances [68,69,70,71,72], but to commercialize this technology, these noble materials should be replaced by more common materials. More studies, such as increasing active surface area using nanotechnology [73,74,75,76,77,78,79], light trapping [80,81,82], Z- scheme photocatalyst [83,84,85,86,87], are required to overcome the current challenges and enhance the efficiency of this technology for commercial use.
Plasmonic nanomaterials can attain frequency-tunable localized-surface plasmon resonances (LSPR) at which the surface electric field is substantially intensified promoting the photocatalytic kinetics reaction rate [88,89]. Plasmonic nanostructures of noble metals (Rh, Ru, Ag, Pd, Pt, and Au) with different photocatalyst materials including TiO2, NiO, GaN:ZnO, have been demonstrated for effective H2 production through water splitting [90,91,92,93,94]. Although there has been considerable advancement in heterojunction photocatalysts, there are serious challenges that need to be addressed to achieve throughput heterojunction photocatalysts at the industrial scale for functional and cost-competitive applications. Moreover, the structure, morphology, porosity, surface conditions, contact interface, and crystallinity of substances need to be accurately controlled to attain high-efficiency heterojunction photocatalysts [66].
Another important scheme for H2 production is power-to-gas (P2G) technology. It uses the surplus renewable electric energy to produce H2 mainly alkaline electrolysis, proton-exchange membrane (PEM) electrolysis, and solid-oxide electrolysis cells (SOECs) [95]. The produced H2 gas can be directly feed into the H2 fuel cells for electric vehicle use or other H2 infrastructures [96]. Alternatively, the H2 can be fed into an NG infrastructure which displaces it and helps to reduce carbon emissions [97]. P2G can also be used to improve the conversion rate of CO2 to biomethane through microbial methanation [98,99]. The overall efficiency for the H2 production in the power-to-gas method is around 60% which is still lower than its competitive storage technologies [100]. However, recent studies have promised 70–80% efficiency using reversible and pressurized reversible solid-oxide cell technologies [101,102]. With the expected improvements, P2G technology is considered to be a promising transition to greener energy and effective use of surplus electricity.
RE and green H2 sources are important for the development of a zero-carbon energy eco-system. To achieve this goal, technologies such as electrolysis, photocatalysis, photoelectrochemical water splitting, and nuclear-based H2 production are being developed, as shown in Table 1. However, methane reforming is still the most commonly used commercially available H2 production technology. It is attractive due to its high efficiency and low-cost operation. Low-carbon emission technologies such as pyrolysis and nuclear-based H2 production are almost ready to be commercially used. However, studies are ongoing to develop technologies where fossil-fuel independent sources (water, sunlight, etc.) are being used. They still suffer from low efficiency, but more work is being done to take these technologies to the level of commercial use.

6. H2 Storage

Although H2 has the highest energy per mass density (about 120 MJ/kg) among all fuels, its low mass density at STP results in low energy per volume density, which necessities the development of sophisticated storage methods (Figure 2) [62]. Storage is a vital issue for the safety, transportation, and utilization aspects of an H2 energy system in stationary, mobile, and off-grid applications. A typical H2-based off-grid system comprises an electrolyzer, a pressurized tank and a FC (to produce electricity from H2 in intermittency periods of REs). An example of this approach was built in the Froan islands (Norway) which well represents typical insular microgrid environments.
Studies show that energy storage devices are the key elements to reduce the dependency on fossil fuels. Due to its affordable prolonged storage capabilities, H2 storage is particularly necessary in off-grid locations to increase the RE penetration, since it prevents the battery and RE technologies becoming bulky or overpriced. The levelized cost of energy (LCOE) for the most affordable option, which contains batteries and H2, is 0.41 €/kWh. This number is around 35% less expensive than the system’s LCOE if it just used batteries to store energy [103].
H2 can be stored either by physical-based methods or material-based methods (Figure 5) [104,105,106,107,108].

6.1. Physical-Based H2 Storage

Pressurized H2 gas is the most established H2 storage method for H2 applications. Since the mass density of H2 is very low (0.089 kg/m3), its storage necessitates high pressure or extremely low temperatures [58]. The current FC applications require pressurization of H2 at the 35–70 mPa level. It was estimated that pressuring would affect 11–13% of the hydrogen energy content negatively [109]. The conventional pressurized H2 storage tanks for commercial applications are made using steel and aluminum materials.
For long-term or large-quantity storage needs, large underground geological formations, such as salt caverns, can be used for compressed H2. Such natural structures are exceptionally impermeable even under high pressures and effectively prevent leakage. A remarkable recent example for underground storage is the HyUnder project in Romania, which is aimed at the storage of enough H2 in salt caverns for potential use in the transportation sector and chemical manufacturing [62]. Storage of H2 in underground salt caverns is a highly scalable and adaptable method. It is much more efficient compared to compressed air energy storage option because hydrogen has a higher energy density. Numerical studies demonstrated that these storage systems could provide emergency backup power for almost seven times longer (57.7 days) than a compressed-air energy system (8.7 days) [110].
Its volumetric mass and energy densities allow H2 to be stored and transported in liquid form in super-insulated containers especially as a way of transmission of large amounts to far points for short term storages to avoid boil-off loss issues. It is an important bridge between rich renewable energy sources and end-users. Liquid H2 has the density of 71 g/L at its boiling point of −253 °C, which is approximately 1.8 times the density of H2 pressurized up to 70 mPa at 15 °C. Due to its low boiling temperature, liquid H2 technology requires cooling down to very cryogenic temperatures (−253 °C) which consume about 30% of its total energy content. Since they are more compact and lighter, cryogenic pressure vessels are considered to be safer than compressed H2 vessels. Liquid H2 storage systems are generally used within a short time (automotive, and air and space applications) [111,112,113].

6.2. Material-Based H2 Storage

In general, an ideal solid-state H2 storage material should be able to absorb a high weight percentage and a high volume of H2, and possess high H2 absorption and desorption speeds at—or close to—STP. It should be inexpensive, resistant to poisoning by trace impurities, have good thermal conductivity in charged and uncharged conditions, require low energy for preparation, be safe and reusable on exposure to air, and have the ability to be regenerated and be readily recycled [54]. The most promising solid-state storage systems for both mobile and stationary applications are material-based. Some of the featured ones among material-based storage techniques are metal hydrides, chemical carriers, and adsorbents.
Metal hydrides are produced by the reaction of H2 with different metals, making solid-state storage under moderate temperature and pressure possible. Hence, they have a unique safety advantage compared to liquid and gaseous storage systems. In metal hydride tanks, H2 molecules bond into metal typically at 3–30 atm pressure, which is much lower than the one in compressed gas tanks. Recently, hydrides of light metals such as Al, Mg, Li, Na, and B have been studied for solid-state storage materials for H2. They present great potential for high safety, volumetric energy density, low-pressure equipment, reversibility of hydration/dehydration, and have low energy requirements. They function like a ‘sponge’ for absorbing and ‘squeezing out’ H2 by small changes in temperature and pressure [114,115]. Aluminum hydride (AlH3) is a covalently bonded tri-hydride with a high volumetric (148 kg.m3) and gravimetric (10.1 wt%) hydrogen capacity. It offers some significant technical difficulties since it is often created by reacting Al with H2 at extremely high hydrogen pressure, yet it decomposes quickly into Al and H2 at relatively low temperatures, demonstrating good hydrogen desorption kinetics at room temperature. AlH2 is thus one of the most promising possibilities for materials for high-capacity hydrogen storage. For hydrogen storage, its synthesis, regeneration, dehydrogenation thermodynamics and kinetics, and techniques for enhancing reversibility are all promising [116]. High percentage of H2 retention is the most important factor for mobile utilizations. Complex hydrides have a greater capability of absorption than intermetallic ones, especially Mg(BH4)2 with its high rate of 14.8 wt.% [104]. A particular member of the metal borohydrides, the nanoporous hydride γ-Mg(BH4)2 demonstrated the capability of storing 17.4 wt.% H2 at 105 bar and −143 °C [117]. The H2 storage density of magnesium hydride (6.5H atoms/cm3) is higher than H2 gas (0.99H atoms/cm3) or liquid H2 (4.2H atoms/cm3) proving that H2 storage in metal hydride is volume-efficient and safe for onboard applications [62,118,119,120].
However, metal hydrides have not yet met all the requirements of H2 energy systems. The challenges that need to be addressed include high weight and low H2 storage capacity of low-temperature metal hydrides, and slow kinetics and undesirable high operation temperatures for lightweight metal hydrides with higher H2 storage capacities [58,62]. Metal hydride tanks can be easily used in stationary applications. In the future, it will be possible to develop lighter materials to incorporate them in mobile applications, and to realize the cost reduction of all H2 storage materials.
Chemical H2 carriers include organic hydrocarbon liquids and inorganic materials such as ammonia and ammonia borane [121]. Liquid-phase hydrogen carriers offer the advantage of using existing petrochemical infrastructure for production, storage, and delivery, e.g., storage tanks and tanker trucks. However, to assess their feasibility more accurately toxicity, material compatibility, lifecycle cost and emission characteristics should be considered.
Hydrogen carriers can be broadly classified as one-way and two-way carriers. One-way carriers include NH3 with a hydrogen content of 109 g/L and 17.6 wt.% and methanol with a hydrogen content of 99 g/L and 12.6 wt.% at 15 °C and 10 bar. These carriers release byproduct gases, such as CO2 and N2, in addition to H2, which would necessitate separation and purification steps for practical applications. Technically, one-way carriers can be regenerated from the products. For example, ammonia can be regenerated from N2 and H2 through the Haber–Bosch process [104,122]. The use of carbon-based one-way carriers such as methanol and formic acid results in the emission of CO2, which requires the integration of CCS modules into the decomposition system.
Two-way carriers, on the other hand, form a byproduct that can be used for regeneration after hydrogen is released. A perfect two-way carrier must be rehydrated for reuse effectively and affordably. For instance, toluene can be hydrogenated to create methylcyclohexane, which can then be dehydrogenated to extract and store the hydrogen. Through catalytic formate/bicarbonate reaction mechanism, formate salts are another possible two-way carrier that may produce and collect hydrogen at appropriate pressures and temperatures. They can be utilized as solids or as aqueous media. H2 storage in organic compounds was not thought to be viable as low-temperature and reversible hydrogen extraction processes are usually regarded as impractical. However, the development of new and efficient catalytic processes can potentially change this assumption [104,123].
In general, chemical carriers have the advantages of relatively dense H2 storage and transport at near-ambient temperatures and pressures. Moreover, it is possible to continuous high-pressure (>120 mPa) H2 production by the selective decomposition of certain chemical carriers with very low level of carbon monoxide release. As the pressure is enough to feed FCVs (70 mPa) without any compression system, it can substantially reduce the H2 infrastructure costs [124].
Nanostorage based on adsorption processes is one of the promising solutions, particularly for mobile systems. Weight and size restrictions for onboard applications require H2 to be stored in small and light vessels, which are supposed to have high gravimetric and volumetric storage capacities. Adsorption to porous and high surface area nanomaterials like carbon nanotubes, metal organic frameworks, activated carbon, spherical carbon fullerenes and metal-decorated carbons are very promising for H2 storage in onboard systems.
Nanomaterials have increased the surface area, shorter diffusion lengths and multiplied grain boundary atoms benefiting thermal-mechanical stability and H2 sorption kinetics. For example, the magnesium hydride + mischmetal nanostructured composite provides decreased grain sizes, an increased free and porous surface, micro-strains and defects. These defects facilitate the diffusion of H2 into the nanostructure by decreasing the diffusion activation energy [125]. Nanomaterials can also be produced by constricting the active material within porous scaffold hosts such as carbon-based materials, zeolites, metal–organic frameworks, and mesoporous silicas and oxides. These hosts are light, have a very large surface space (to facilitate higher contents), extremely porous, are non-reactive, cheap, abundant, and have a suitable purity. The large surface areas of carbon-based materials (nanorods, nanofibers, nanowires, nanotubes, and nanoribbons) that improve bulk adsorption/desorption temperature, physisorption kinetics, surface interactions, and H2 dissociation have attracted attention [62].
In addition to carbon-based nanostructures, there are other nanostructures offering strong potential for H2 storage. Boron fullerenes and boron nitride nanostructures are two prominent families of these materials, with similar geometries to carbon nanostructures but with different properties [126]. Studies on calcium-decorated boron fullerenes using density functional theory revealed that the decorating calcium atoms can attract five H2 molecules each. With favorable binding energies of 0.12–0.40 eV/H2, a B80 system adorned with 12 calcium atoms exhibited a gravimetric density of 8.2 wt% hydrogen. Similarly, a boron nanotube decorated with Ca showed hydrogen storage capacity of 7.6 weight percent with binding energies of 0.10–0.30 eV/H2. Most notably, ambient environmental conditions were suitable for the adsorption and desorption of these H2 molecules [127].
Different H2 storage technologies are summarized in Table 2, along with their key parameters and important applications.

7. H2 Distribution

To distribute H2 produced in remote RE-rich points to dense population regions or electricity generation plants, a wide infrastructure and a grid of distribution and H2 refueling stations are required. This depends on many parameters, such as distance, amount of H2, large scale storage possibilities, and final use. It should be noted that in some cases, insular environments compel and allow for producing and utilizing low-cost H2. Worldwide established NG infrastructure and pipelines have the flexibility to be converted to H2 transmission systems. In this scenario “fugitive emissions of H2 infrastructures, similar to those of NG” are possible [109].
Injecting H2 into the NG distribution networks is technically feasible. Based on the individual network and/or location, mixes of H2 up to 20% by volume may be feasible without the need for infrastructure conversion, according to the EU’s H2 agenda. To assess the feasibility of using high-pressure pipelines for H2 transmission at a broader scale, the impact of H2 depending on the materials that are utilized, and the pressure levels, need to be investigated further. The goals of the HyDeploy (UK) and Hyblend (US) programs are to comprehend how greater H2 concentrations could affect various gas grid components as well as the potential effects on piping and pipeline materials [47]. Some existing pipelines already have properties to meet requirements to be used for H2 distribution. For example, a 12 km pipeline formerly used for transporting NG in the Netherlands has been transformed for transportation of 100% H2 in 2018. The NortH2 project in the same country aims to produce and distribute 800,000 tons of H2 through the NG infrastructure. Another sign of the growing momentum for repurposing NG pipelines in Europe is the European H2 Backbone (EHB) effort. To repurpose a 39,700 km H2 pipeline system that is expanding in 21 nations, the EHB now engages 23 gas infrastructure corporations from 21 countries [47]. Results of these initiatives and pilot projects are expected to build a strong knowledge base to develop larger scale H2 distribution infrastructure in the near future.

8. H2 Utilization

In line with the worldwide efforts and developments aiming to attain climate goals on the one hand, and an international energy crisis on the other hand, almost all industries and transport sectors evaluate H2 alternatives more seriously than in the past and tend to establish their H2 utilization systems. The H2 gas turbine is one such system, albeit with some challenges. After different reconversion technologies were evaluated, there is now an H2 gas turbine technology which runs on 100% green H2 and makes it possible to convert conventional NG turbines to those of H2. Moreover, H2 gas turbines may have a synergistic benefit from the existing H2 infrastructures built for other industrial applications, such as those in the transportation sector and in fossil-fuel steel production.
As a part of its 2050 greenhouse gas-neutral status goal, Germany plans to feed all H2 gas turbines with 100% green H2 generated by electrolysis. The plan’s modeling showed that H2-fueled gas turbines can be competitive under severe CO2 emission caps (as in the modeling for the years 2040 and 2050) and under conditions where variable RE power is widely used [128,129]. In fact, the town of Lingen in northern Germany will soon have a 34 MW H2 gas turbine power station that runs on 100% green H2, offering a carbon-free solution to electricity generation on an industrial scale, with an efficiency of 40.3% [130].
On the other hand, H2 gas turbines come with their own challenges. Hydrogen flame propagation velocity is up to seven times higher than NG, which may lead to flame instability, undesirable pressure fluctuations, and mechanical stress on the parts of the combustor, all of which require careful design optimization. Moreover, H2 burns at a higher temperature compared to NG, which results in up to three times more nitrogen oxides (NOx) than NG. Being a toxic air pollutant, NOx needs to be reduced/eliminated through careful turbine design and integrated capturing systems.
In another field of H2 utilization, FCs convert H2 into electricity with an efficiency more than twice, the efficiency of ICEs. H2 FCs engines operate in transportation with an efficiency of up to 65%, compared to 25% for conventional petroleum car engines. The overall efficiency of FCs even exceeds 85% when the generated heat is used in combined heat and power systems. This offers an enormous advantage for FCs, considering that the typical efficiency of a modern coal-fired power plant is ~45% [2,52]. In 2018, Germany started the world’s first H2 FC powered commercial train service that can run for about 1000 km on a single tank of hydrogen, similar to the range of diesel trains [131].

9. Global Zero-Carbon Green H2 Economy

In 2022, about 80% of global energy demand is met by oil, NG, and coal. The energy sector is responsible for about three-quarters of GHG emissions, with electricity generation alone accounting for 40% of it. Another 60% or so is emitted by fossils fuels used in industry, transportation, and heating [37,132]. The research shows that a safe future below 1.5 °C global warming requires the world to cut 30 gigatons of GHG emissions annually by 2030 [133]. Within this effort, H2 is particularly appealing for the energy sector due to its broad usage area. It may serve as an alternative engine fuel, a means of generating and storing electricity and heat. The heating applications already constitute the third largest consumer of H2 today [34].
The global climate goals require significant levels of new investment to successfully commercialize and scale a global green H2 industry. It is estimated that global annual investments of between US $20 to $25 billion are needed for a total investment of about $280 billion by 2030 to provide up to 18% of the world’s energy demand with H2 by 2050 [15]. Given the already available industrial infrastructure and compatible distribution networks, for H2, it is not a feedstock and technology problem, but a cost and time problem. H2 gives the possibility to develop multiple alternatives both in its integration with RE in terms of production, and in its post-production processes, provided that its storage and distribution network costs are subsidized for a while [37].

10. Conclusions

Current climate action plans of governments are crucial to achieve the Paris Agreement target (for which, in November 2022 a commitment was repeated to keep the plans alive at the COP27 in Sharm El Sheikh, Egypt) of holding global average temperature increase to a maximum of 1.5 °C. There are some pushbacks from energy companies regarding a reduction in fossil-fuel production. A clean and sustainable alternative is needed to combat the idea that the climate is no longer a priority. As a clean, renewable energy carrier, H2 produced via electrolysis, photocatalysis, and photoelectrochemical splitting of ubiquitously available water of which electricity comes from clean REs (wind, solar, wave, tidal energies, and hydro, and geothermal power plants) is the most promising synthetic fuel in the transition period to zero carbon. Although electrolytic H2 alone cannot meet the demand of H2 for the world, at least in the upcoming years, green H2 can compensate for the mismatch between production of and the demand for RE. Today, H2 is already used in almost all types of vehicles and industry. It can be used in power plants to generate electricity. FCs can substitute conventional engines. Green H2 systems present a considerable integration capability with the existing industrial use of H2 infrastructure, as well as with clean RE energy sources and fossil-fuel infrastructures. RE sources used in clean electricity production are vital for H2 sustainability in addition to their other uses. Their intermittency from spatiotemporal instabilities can be balanced by storage in the form of H2 produced by clean energy through electrolysis. Such an integrated H2 eco-system and economy require a cost-effective, large-scale H2 infrastructure. The rapidly decreasing cost of electricity generated from REs further encourages investment in the widespread production and utilization of green H2 by the middle of the 21st century. Leading up to that, low carbon emission H2 generation processes such as nuclear-produced H2, pyrolysis, and steam reforming with CCS using fossil-fuel as the source will play a vital role in the upcoming years. These processes have the ability to keep carbon emissions in check, in some cases (nuclear) producing almost zero-carbon emissions. These technologies will help meet the demand of H2 in the transition period while green H2 generation technologies mature to meet global demand. Today, for H2, it is not a question of competitiveness but a global will and determination to invest more in electrolyzers, clean REs and other infrastructure components in the face of global climate change.

Author Contributions

All authors have contributed to this paper equally. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Crippa, M.; Guizzardi, D.; Banja, M.; Solazzo, E.; Muntean, M.; Schaaf, E.; Pagani, F.; Monforti-Ferrario, F.; Olivier, J.; Quadrelli, R.; et al. CO2 Emissions of All World Countries—2022 Report; EUR 31182 EN; Publications Office of the European Union: Luxembourg, 2022. [Google Scholar]
  2. IEA. The Future of H2: Report prepared by the IEA for the G20, Japan; IEA: Paris, France, 2019. [Google Scholar]
  3. Ari, A.; Arregui, N.; Black, S.; Celasun, O.; Iakova, D.; Mineshima, A.; Mylonas, V.; Parry, I.; Teodoru, I.; Zhunussova, K. Surging Energy Prices in Europe in the Aftermath of the War: How to Support the Vulnerable and Speed up the Transition Away from Fossil Fuels; International Monetary Fund: Washington, DC, USA, 2022. [Google Scholar]
  4. Öztürk, Y. Reflections of the Russia-Ukraine War on the Energy Sector. Synergy 2022. Available online: https://www.bilkenteprc.com/post/reflections-of-the-russia-ukraine-war-on-the-energy-sector-yaren-%C3%B6zt%C3%BCrk (accessed on 29 August 2022).
  5. Resource Trade Earth Dashboard. Available online: https://resourcetrade.earth/ (accessed on 21 November 2022).
  6. Carroll, S.G. Biofuels’ impact on food security debate resurfaces amid Ukraine war. Euractiv, 8 April 2022. [Google Scholar]
  7. Boucher, H. Guest view: Global hunger fight means no biofuel. Reuters, 6 June 2022. [Google Scholar]
  8. Hosseini, S.E. Transition away from fossil fuels toward renewables: Lessons from Russia-Ukraine crisis. Future Energy 2022, 1, 2–5. [Google Scholar] [CrossRef]
  9. Taghizadeh-Hesary, F.; Li, Y.; Rasoulinezhad, E.; Mortha, A.; Long, Y.; Lan, Y.; Zhang, Z.; Li, N.; Zhao, X.; Wang, Y. Green finance and the economic feasibility of hydrogen projects. Int. J. Hydrogen Energy 2022, 47, 24511–24522. [Google Scholar] [CrossRef]
  10. 14th Five-year Plan for Renewable Energy Development. Available online: https://chinaenergyportal.org/en/14th-five-year-plan-for-renewable-energy-development/ (accessed on 25 December 2022).
  11. Samsun, R.C.; Rex, M.; Antoni, L.; Stolten, D. Deployment of fuel cell vehicles and hydrogen refueling station infrastructure: A global overview and perspectives. Energies 2022, 15, 4975. [Google Scholar] [CrossRef]
  12. Wood, J. Which Countries Could Become the World’s Hydrogen Superpowers? Available online: https://www.weforum.org/agenda/2022/02/clean-hydrogen-energy-low-carbon-superpowers/ (accessed on 21 November 2022).
  13. Hosseini, S.E. Hydrogen has found its way to become the fuel of the future. Future Energy 2022, 1, 11–12. [Google Scholar] [CrossRef]
  14. Office of Energy Efficiency & Renewable Energy: Hydrogen Storage. Available online: https://www.energy.gov/eere/fuelcells/hydrogen-storage (accessed on 25 November 2022).
  15. Dong, Z.Y.; Yang, J.; Yu, L.; Daiyan, R.; Amal, R. A green hydrogen credit framework for international green hydrogen trading towards a carbon neutral future. Int. J. Hydrogen Energy 2022, 47, 728–734. [Google Scholar] [CrossRef]
  16. IRENA. Renewable Power Generation Costs in 2021; International Renewable Energy Agency (IRENA): Abu Dhabi, United Arab Emirates, 2022. [Google Scholar]
  17. Baldino, C.; O’Malley, J.; Searle, S.; Zhou, Y.; Christensen, A. Hydrogen for heating? Decarbonization options for households in the United Kingdom in 2050. Int. Counc. Clean Transp. 2020. Available online: https://theicct.org/publication/hydrogen-for-heating-decarbonization-options-for-households-in-the-united-kingdom-in-2050/ (accessed on 25 December 2022).
  18. Longden, T.; Beck, F.J.; Jotzo, F.; Andrews, R.; Prasad, M. ‘Clean’ hydrogen?—Comparing the emissions and costs of fossil fuel versus renewable electricity based hydrogen. Appl. Energy 2022, 306, 118145. [Google Scholar] [CrossRef]
  19. Agency, I.R.E. Green Hydrogen Cost Reduction: Scaling up Electrolysers to Meet the 1.5 °C Climate Goal; International Renewable Energy Agency: Abu Dhabi, United Arab Emirates, 2020. [Google Scholar]
  20. Stöckl, F.; Schill, W.-P.; Zerrahn, A. Optimal supply chains and power sector benefits of green hydrogen. Sci. Rep. 2021, 11, 14191. [Google Scholar] [CrossRef]
  21. Mao, S.; Basma, H.; Ragon, P.-L.; Zhou, Y.; Rodríguez, F. Total Cost of Ownership for Heavy Trucks in China: Battery Electric, Fuel Cell, and Diesel Trucks; National Academy of Sciences: Washington, DC, USA, 2021. [Google Scholar]
  22. Zhou, Y.; Searle, S.; Baldino, C. Cost of Renewable Hydrogen Produced Onsite at Hydrogen Refueling Station in Europe; The International Council on Clean Transportation: San Francisco, CA, USA, 2022. [Google Scholar]
  23. Siddiqui, O.; Dincer, I. Optimization of a new renewable energy system for producing electricity, hydrogen and ammonia. Sustain. Energy Technol. Assess. 2021, 44, 101023. [Google Scholar] [CrossRef]
  24. Abdin, Z.; Webb, C.; Gray, E. Solar hydrogen hybrid energy systems for off-grid electricity supply: A critical review. Renew. Sustain. Energy Rev. 2015, 52, 1791–1808. [Google Scholar] [CrossRef]
  25. Yue, M.; Lambert, H.; Pahon, E.; Roche, R.; Jemei, S.; Hissel, D. Hydrogen energy systems: A critical review of technologies, applications, trends and challenges. Renew. Sustain. Energy Rev. 2021, 146, 111180. [Google Scholar] [CrossRef]
  26. Avargani, V.M.; Zendehboudi, S.; Saady, N.M.C.; Dusseault, M.B. A comprehensive review on hydrogen production and utilization in North America: Prospects and challenges. Energy Convers. Manag. 2022, 269, 115927. [Google Scholar]
  27. Sorensen, B. Hydrogen and Fuel Cells: Emerging Technologies and Applications; Academic Press: New York, NY, USA, 2011. [Google Scholar]
  28. Demirbas, M.F. Hydrogen from various biomass species via pyrolysis and steam gasification processes. Energy Sources Part A 2006, 28, 245–252. [Google Scholar] [CrossRef]
  29. Singh, V.; Buelens, L.C.; Poelman, H.; Saeys, M.; Marin, G.B.; Galvita, V.V. Intensifying blue hydrogen production by in situ CO2 utilisation. J. CO2 Util. 2022, 61, 102014. [Google Scholar] [CrossRef]
  30. Asadullah, M.; Ito, S.-I.; Kunimori, K.; Yamada, M.; Tomishige, K. Energy efficient production of hydrogen and syngas from biomass: Development of low-temperature catalytic process for cellulose gasification. Environ. Sci. Technol. 2002, 36, 4476–4481. [Google Scholar] [CrossRef]
  31. Holladay, J.D.; Hu, J.; King, D.L.; Wang, Y. An overview of hydrogen production technologies. Catal. Today 2009, 139, 244–260. [Google Scholar] [CrossRef]
  32. Birol, F. The future of hydrogen: Seizing today’s opportunities. IEA Rep. Prep. G 2019, 20. Available online: https://0-read-oecd--ilibrary-org.brum.beds.ac.uk/energy/the-future-of-hydrogen_1e0514c4-en (accessed on 25 December 2022).
  33. Watson, N.; Donovan, J. IAEA Modelling Shows High NG Prices Shift Optimal H2 Production to Nuclear Energy. IAEA Department of Nuclear Energy, 28 October 2021. [Google Scholar]
  34. Sönnichsen, N. Global hydrogen production outlook by type 2015–2050. Available online: https://0-www-statista-com.brum.beds.ac.uk/statistics/859104/hydrogen-production-outlook-worldwide-by-type/ (accessed on 29 August 2022).
  35. Bauer, C.; Treyer, K.; Antonini, C.; Bergerson, J.; Gazzani, M.; Gencer, E.; Gibbins, J.; Mazzotti, M.; McCoy, S.T.; McKenna, R. On the climate impacts of blue hydrogen production. Sustain. Energy Fuels 2022, 6, 66–75. [Google Scholar] [CrossRef]
  36. Howarth, R.W.; Jacobson, M.Z. How green is blue hydrogen? Energy Sci. Eng. 2021, 9, 1676–1687. [Google Scholar] [CrossRef]
  37. Mac Dowell, N.; Sunny, N.; Brandon, N.; Herzog, H.; Ku, A.Y.; Maas, W.; Ramirez, A.; Reiner, D.M.; Sant, G.N.; Shah, N. The hydrogen economy: A pragmatic path forward. Joule 2021, 5, 2524–2529. [Google Scholar] [CrossRef]
  38. Cloete, S.; del Pozo, C.A.; Álvaro, Á.J. System-friendly process design: Optimizing blue hydrogen production for future energy systems. Energy 2022, 259, 124954. [Google Scholar] [CrossRef]
  39. Schneider, S.; Bajohr, S.; Graf, F.; Kolb, T. State of the art of hydrogen production via pyrolysis of natural gas. ChemBioEng Rev. 2020, 7, 150–158. [Google Scholar] [CrossRef]
  40. Ashton, L.; Fast-Tracking Nuclear Hydrogen: IAEA to Develop Roadmap for Commercial Deployment. International Atomic Energy Agency (IAEA) 5 May 2022. Available online: https://www.iaea.org/newscenter/news/fast-tracking-nuclear-hydrogen-iaea-to-develop-roadmap-for-commercial-deployment (accessed on 29 August 2022).
  41. IEA. H2 Production and Storage, R&D Priorities and Gaps; International Energy Agency (IEA): Paris, France, 2006. [Google Scholar]
  42. Ashton, L. IAEA Ministerial Conference to Highlight Nuclear Power’s Role in Achieving Net Zero and Energy Security; International Atomic Energy Agency (IAEA): Vienna, Austria, 2022. [Google Scholar]
  43. US Department of Energy. Energy Department Announces up to $3.5M for Nuclear-Compatible Hydrogen Production Office of Energy Efficiency & Renewable Energy; US Department of Energy: Washington, DC, USA, 2018.
  44. Watson, N.; Constantin, A. IAEA Event Showcases Progress, Innovations in Nuclear Hydrogen for a Clean Energy Transition; International Atomic Energy Agency (IAEA): Vienna, Austria, 2021. [Google Scholar]
  45. BloombergNEF. H2 Economy Outlook; Bloomberg Finance L.P. 2020. Available online: https://data.bloomberglp.com/professional/sites/24/BNEF-Hydrogen-Economy-Outlook-Key-Messages-30-Mar-2020.pdf (accessed on 29 August 2022).
  46. European Commission. H2 Energy and FCSs: A Vision of Our Future; European Commission Directorate-General for Research: Brussels, Belgium, 2003; Available online: https://inis.iaea.org/collection/NCLCollectionStore/_Public/37/121/37121708.pdf (accessed on 29 August 2022).
  47. Clean H2 Joint Undertaking, E.C. Strategic Research and Innovation Agenda 2021–2027, Annex to GB decision no. CleanHydrogen-GB-2022-02. 25 February 2022. Available online: https://www.clean-hydrogen.europa.eu/system/files/2022-02/Clean%20Hydrogen%20JU%20SRIA%20-%20approved%20by%20GB%20-%20clean%20for%20publication%20%28ID%2013246486%29.pdf (accessed on 29 August 2022).
  48. Zorpette, G. 2022—The Year the Hydrogen Economy Launched? IEEE Spectr. 17 August 2022. Available online: https://0-spectrum-ieee-org.brum.beds.ac.uk/hydrogen-economy-inflation-reduction-act (accessed on 29 August 2022).
  49. Huang, Y.-S.; Liu, S.-J. Chinese green hydrogen production potential development: A provincial case study. IEEE Access 2020, 8, 171968–171976. [Google Scholar] [CrossRef]
  50. Tang, O.; Rehme, J.; Cerin, P. Levelized cost of hydrogen for refueling stations with solar PV and wind in Sweden: On-grid or off-grid? Energy 2022, 241, 122906. [Google Scholar] [CrossRef]
  51. Ni, N.; Qie, B.; Du, S.; Sang, Z.; Wang, Q.; Meng, C.; Tong, Y. A novel all-solid-state S-scheme in CdS/ZnTHPP binary nanosystem for hydrogen evolution. Int. J. Hydrogen Energy 2022, 47, 13044–13053. [Google Scholar] [CrossRef]
  52. Dutton, G. Hydrogen Energy Technology; Tyndall Centre for Climate Change Research: Norwich, UK, 2002; Volume 31. [Google Scholar]
  53. Ewan, B.; Allen, R. A figure of merit assessment of the routes to hydrogen. Int. J. Hydrogen Energy 2005, 30, 809–819. [Google Scholar] [CrossRef]
  54. Noyan, Ö.F. Some approach to possible atmospheric impacts of a hydrogen energy system in the light of the geological past and present-day. Int. J. Hydrogen Energy 2011, 36, 11216–11228. [Google Scholar] [CrossRef]
  55. Kumar, S.S.; Himabindu, V. Hydrogen production by PEM water electrolysis—A review. Mater. Sci. Energy Technol. 2019, 2, 442–454. [Google Scholar]
  56. Younas, M.; Shafique, S.; Hafeez, A.; Javed, F.; Rehman, F. An overview of hydrogen production: Current status, potential, and challenges. Fuel 2022, 316, 123317. [Google Scholar] [CrossRef]
  57. Naito, T.; Shinagawa, T.; Nishimoto, T.; Takanabe, K. Gas Crossover Regulation by Porosity—Controlled Glass Sheet Achieves Pure Hydrogen Production by Buffered Water Electrolysis at Neutral pH. ChemSusChem 2022, 15, e202102294. [Google Scholar] [CrossRef]
  58. Bodner, M.; Hofer, A.; Hacker, V. H2 generation from alkaline electrolyzer. Wiley Interdiscip. Rev. Energy Environ. 2015, 4, 365–381. [Google Scholar]
  59. Pinsky, R.; Sabharwall, P.; Hartvigsen, J.; O’Brien, J. Comparative review of hydrogen production technologies for nuclear hybrid energy systems. Prog. Nucl. Energy 2020, 123, 103317. [Google Scholar] [CrossRef]
  60. Chu, J. MIT’s MOXIE Experiment Reliably Produces Oxygen on Mars; Massachusetts Institute of Technology: Cambridge, MA, USA, 2022. [Google Scholar]
  61. Guo, J.; Zhang, Y.; Zavabeti, A. Hydrogen production from the air. Nat. Commun. 2022, 13, 5046. [Google Scholar] [CrossRef] [PubMed]
  62. Abe, J.O.; Popoola, A.; Ajenifuja, E.; Popoola, O. Hydrogen energy, economy and storage: Review and recommendation. Int. J. Hydrogen Energy 2019, 44, 15072–15086. [Google Scholar] [CrossRef]
  63. Burton, N.; Padilla, R.; Rose, A.; Habibullah, H. Increasing the efficiency of hydrogen production from solar powered water electrolysis. Renew. Sustain. Energy Rev. 2021, 135, 110255. [Google Scholar] [CrossRef]
  64. Arunachalam, P.; Al Mayouf, A.M. Photoelectrochemical water splitting. In Noble Metal-Metal Oxide Hybrid Nanoparticles; Elsevier: Amsterdam, The Netherlands, 2019; pp. 585–606. [Google Scholar]
  65. Wu, H.; Tan, H.L.; Toe, C.Y.; Scott, J.; Wang, L.; Amal, R.; Ng, Y.H. Photocatalytic and photoelectrochemical systems: Similarities and differences. Adv. Mater. 2020, 32, 1904717. [Google Scholar] [CrossRef] [PubMed]
  66. Bahadoran, A.; Liu, Q.; Ramakrishna, S.; Sadeghi, B.; De Castro, M.M.; Cavaliere, P.D. Hydrogen Production as a Clean Energy Carrier through Heterojunction Semiconductors for Environmental Remediation. Energies 2022, 15, 3222. [Google Scholar] [CrossRef]
  67. Zhou, H.; Wen, Z.; Liu, J.; Ke, J.; Duan, X.; Wang, S. Z-scheme plasmonic Ag decorated WO3/Bi2WO6 hybrids for enhanced photocatalytic abatement of chlorinated-VOCs under solar light irradiation. Appl. Catal. B Environ. 2019, 242, 76–84. [Google Scholar] [CrossRef]
  68. Wang, Q.; Hisatomi, T.; Jia, Q.; Tokudome, H.; Zhong, M.; Wang, C.; Pan, Z.; Takata, T.; Nakabayashi, M.; Shibata, N. Scalable water splitting on particulate photocatalyst sheets with a solar-to-hydrogen energy conversion efficiency exceeding 1%. Nat. Mater. 2016, 15, 611–615. [Google Scholar] [CrossRef]
  69. Fujito, H.; Kunioku, H.; Kato, D.; Suzuki, H.; Higashi, M.; Kageyama, H.; Abe, R. Layered perovskite oxychloride Bi4NbO8Cl: A stable visible light responsive photocatalyst for water splitting. J. Am. Chem. Soc. 2016, 138, 2082–2085. [Google Scholar] [CrossRef] [PubMed]
  70. Chen, S.; Qi, Y.; Hisatomi, T.; Ding, Q.; Asai, T.; Li, Z.; Ma, S.S.K.; Zhang, F.; Domen, K.; Li, C. Efficient visible-light-driven Z-scheme overall water splitting using a MgTa2O6− xNy/TaON heterostructure photocatalyst for H2 evolution. Angew. Chem. 2015, 127, 8618–8621. [Google Scholar] [CrossRef]
  71. Jia, Q.; Iwase, A.; Kudo, A. BiVO4–Ru/SrTiO3: Rh composite Z-scheme photocatalyst for solar water splitting. Chem. Sci. 2014, 5, 1513–1519. [Google Scholar] [CrossRef]
  72. Chowdhury, F.A.; Trudeau, M.L.; Guo, H.; Mi, Z. A photochemical diode artificial photosynthesis system for unassisted high efficiency overall pure water splitting. Nat. Commun. 2018, 9, 1707. [Google Scholar] [CrossRef] [PubMed]
  73. Tan, H.L.; Amal, R.; Ng, Y.H. Exploring the different roles of particle size in photoelectrochemical and photocatalytic water oxidation on BiVO4. ACS Appl. Mater. Interfaces 2016, 8, 28607–28614. [Google Scholar] [CrossRef] [PubMed]
  74. Cheng, B.; Samulski, E.T. Hydrothermal synthesis of one-dimensional ZnO nanostructures with different aspect ratios. Chem. Commun. 2004, 8, 986–987. [Google Scholar] [CrossRef] [PubMed]
  75. Mishra, A.K.; Pradhan, D. Morphology controlled solution-based synthesis of Cu2O crystals for the facets-dependent catalytic reduction of highly toxic aqueous Cr (VI). Cryst. Growth Des. 2016, 16, 3688–3698. [Google Scholar] [CrossRef]
  76. Dai, P.; Xie, J.; Mayer, M.T.; Yang, X.; Zhan, J.; Wang, D. Solar hydrogen generation by silicon nanowires modified with platinum nanoparticle catalysts by atomic layer deposition. Angew. Chem. 2013, 125, 11325–11329. [Google Scholar] [CrossRef]
  77. Xu, Q.C.; Wellia, D.V.; Ng, Y.H.; Amal, R.; Tan, T.T.Y. Synthesis of porous and visible-light absorbing Bi2WO6/TiO2 heterojunction films with improved photoelectrochemical and photocatalytic performances. J. Phys. Chem. C 2011, 115, 7419–7428. [Google Scholar] [CrossRef]
  78. Shim, H.-S.; Shinde, V.R.; Kim, J.W.; Gujar, T.P.; Joo, O.-S.; Kim, H.J.; Kim, W.B. Diameter-tunable CdSe nanotubes from facile solution-based selenization of Cd (OH)2 nanowire bundles for photoelectrochemical cells. Chem. Mater. 2009, 21, 1875–1883. [Google Scholar] [CrossRef]
  79. Lin, X.; Xie, Z.; Su, B.; Zheng, M.; Dai, W.; Hou, Y.; Ding, Z.; Lin, W.; Fang, Y.; Wang, S. Well-defined Co9 S8 cages enable the separation of photoexcited charges to promote visible-light CO2 reduction. Nanoscale 2021, 13, 18070–18076. [Google Scholar] [CrossRef] [PubMed]
  80. Yu, H.; Bai, Y.; Zong, X.; Tang, F.; Lu, G.M.; Wang, L. Cubic CeO2 nanoparticles as mirror-like scattering layers for efficient light harvesting in dye-sensitized solar cells. Chem. Commun. 2012, 48, 7386–7388. [Google Scholar] [CrossRef] [PubMed]
  81. Kim, K.; Moon, J.H. Three-dimensional bicontinuous BiVO4/ZnO photoanodes for high solar water-splitting performance at low bias potential. ACS Appl. Mater. Interfaces 2018, 10, 34238–34244. [Google Scholar] [CrossRef]
  82. Zhang, L.; Reisner, E.; Baumberg, J.J. Al-doped ZnO inverse opal networks as efficient electron collectors in BiVO4 photoanodes for solar water oxidation. Energy Environ. Sci. 2014, 7, 1402–1408. [Google Scholar] [CrossRef]
  83. Zhang, P.; Wang, T.; Chang, X.; Gong, J. Effective charge carrier utilization in photocatalytic conversions. Acc. Chem. Res. 2016, 49, 911–921. [Google Scholar] [CrossRef]
  84. Soldat, J.; Marschall, R.; Wark, M. Improved overall water splitting with barium tantalate mixed oxide composites. Chem. Sci. 2014, 5, 3746–3752. [Google Scholar] [CrossRef] [Green Version]
  85. Yu, W.; Zhang, S.; Chen, J.; Xia, P.; Richter, M.H.; Chen, L.; Xu, W.; Jin, J.; Chen, S.; Peng, T. Biomimetic Z-scheme photocatalyst with a tandem solid-state electron flow catalyzing H 2 evolution. J. Mater. Chem. A 2018, 6, 15668–15674. [Google Scholar] [CrossRef]
  86. Wang, F.; Li, Q.; Xu, D. Recent progress in semiconductor-based nanocomposite photocatalysts for solar-to-chemical energy conversion. Adv. Energy Mater. 2017, 7, 1700529. [Google Scholar] [CrossRef]
  87. Li, X.-H.; Wang, B.-J.; Wang, G.-D.; Ke, S.-H. Blue phosphorene/Sc2 CX2 (X = O, F) van der Waals heterostructures as suitable candidates for water-splitting photocatalysts and solar cells. Sustain. Energy Fuels 2020, 4, 5277–5283. [Google Scholar] [CrossRef]
  88. Gellé, A.; Jin, T.; de la Garza, L.; Price, G.D.; Besteiro, L.V.; Moores, A. Applications of plasmon-enhanced nanocatalysis to organic transformations. Chem. Rev. 2019, 120, 986–1041. [Google Scholar] [CrossRef]
  89. Linic, S.; Chavez, S.; Elias, R. Flow and extraction of energy and charge carriers in hybrid plasmonic nanostructures. Nat. Mater. 2021, 20, 916–924. [Google Scholar] [CrossRef]
  90. Reddy, N.L.; Rao, V.N.; Vijayakumar, M.; Santhosh, R.; Anandan, S.; Karthik, M.; Shankar, M.; Reddy, K.R.; Shetti, N.P.; Nadagouda, M. A review on frontiers in plasmonic nano-photocatalysts for hydrogen production. Int. J. Hydrogen Energy 2019, 44, 10453–10472. [Google Scholar] [CrossRef]
  91. Belessiotis, G.V.; Kontos, A.G. Plasmonic silver (Ag)-based photocatalysts for H2 production and CO2 conversion: Review, analysis and perspectives. Renew. Energy 2022, 195, 497–515. [Google Scholar] [CrossRef]
  92. Ezendam, S.; Herran, M.; Nan, L.; Gruber, C.; Kang, Y.; Grobmeyer, F.; Lin, R.; Gargiulo, J.; Sousa-Castillo, A.; Cortés, E. Hybrid Plasmonic Nanomaterials for Hydrogen Generation and Carbon Dioxide Reduction. ACS Energy Lett. 2022, 7, 778–815. [Google Scholar] [CrossRef] [PubMed]
  93. Kumar, A.; Choudhary, P.; Kumar, A.; Camargo, P.H.; Krishnan, V. Recent advances in plasmonic photocatalysis based on TiO2 and noble metal nanoparticles for energy conversion, environmental remediation, and organic synthesis. Small 2022, 18, 2101638. [Google Scholar] [CrossRef] [PubMed]
  94. Yuan, L.; Zhou, J.; Zhang, M.; Wen, X.; Martirez, J.M.P.; Robatjazi, H.; Zhou, L.; Carter, E.A.; Nordlander, P.; Halas, N.J. Plasmonic Photocatalysis with Chemically and Spatially Specific Antenna–Dual Reactor Complexes. ACS Nano 2022, 16, 17365–17375. [Google Scholar] [CrossRef]
  95. Liu, W.; Wen, F.; Xue, Y. Power-to-gas technology in energy systems: Current status and prospects of potential operation strategies. J. Mod. Power Syst. Clean Energy 2017, 5, 439–450. [Google Scholar] [CrossRef]
  96. Eberle, U.; Müller, B.; Von Helmolt, R. Fuel cell electric vehicles and hydrogen infrastructure: Status 2012. Energy Environ. Sci. 2012, 5, 8780–8798. [Google Scholar] [CrossRef]
  97. Melaina, M.W.; Antonia, O.; Penev, M. Blending Hydrogen into Natural Gas Pipeline Networks: A Review of Key Issues; US Department of Energy: Washington, DC, USA, 2013.
  98. Yates, M.D.; Siegert, M.; Logan, B.E. Hydrogen evolution catalyzed by viable and non-viable cells on biocathodes. Int. J. Hydrogen Energy 2014, 39, 16841–16851. [Google Scholar] [CrossRef]
  99. Deutzmann, J.S.; Sahin, M.; Spormann, A.M. Extracellular enzymes facilitate electron uptake in biocorrosion and bioelectrosynthesis. MBio 2015, 6, e00496-15. [Google Scholar] [CrossRef]
  100. Quaschning, V. Renewable Energy Systems. Technology-Calculation-Simulation. 8. upd. and enl. ed.; Regenerative Energiesysteme. Technologie-Berechnung-Simulation; Hanser Fachbuchverlag: Berlin, Germany, 2013. [Google Scholar]
  101. Jensen, S.H.; Graves, C.; Mogensen, M.; Wendel, C.; Braun, R.; Hughes, G.; Gao, Z.; Barnett, S.A. Large-scale electricity storage utilizing reversible solid oxide cells combined with underground storage of CO2 and CH4. Energy Environ. Sci. 2015, 8, 2471–2479. [Google Scholar] [CrossRef]
  102. Butera, G.; Jensen, S.H.; Clausen, L.R. A novel system for large-scale storage of electricity as synthetic natural gas using reversible pressurized solid oxide cells. Energy 2019, 166, 738–754. [Google Scholar] [CrossRef]
  103. Marocco, P.; Ferrero, D.; Lanzini, A.; Santarelli, M. The role of hydrogen in the optimal design of off-grid hybrid renewable energy systems. J. Energy Storage 2022, 46, 103893. [Google Scholar] [CrossRef]
  104. Wieliczko, M.; Stetson, N. Hydrogen technologies for energy storage: A perspective. MRS Energy Sustain. 2020, 7, E41. [Google Scholar] [CrossRef]
  105. Rivard, E.; Trudeau, M.; Zaghib, K. Hydrogen storage for mobility: A review. Materials 2019, 12, 1973. [Google Scholar] [CrossRef] [PubMed]
  106. Hassan, I.; Ramadan, H.S.; Saleh, M.A.; Hissel, D. Hydrogen storage technologies for stationary and mobile applications: Review, analysis and perspectives. Renew. Sustain. Energy Rev. 2021, 149, 111311. [Google Scholar]
  107. Usman, M.R. Hydrogen storage methods: Review and current status. Renew. Sustain. Energy Rev. 2022, 167, 112743. [Google Scholar] [CrossRef]
  108. Tarhan, C.; Çil, M.A. A study on hydrogen, the clean energy of the future: Hydrogen storage methods. J. Energy Storage 2021, 40, 102676. [Google Scholar] [CrossRef]
  109. Pesonen, O.; Alakunnas, T. Energy Storage: A Missing Piece of the Puzzle for the Self-Sufficient Living; Lapland University of Applied Sciences: Rovaniemi, Finland, 2017. [Google Scholar]
  110. Takach, M.; Sarajlić, M.; Peters, D.; Kroener, M.; Schuldt, F.; von Maydell, K. Review of Hydrogen Production Techniques from Water Using Renewable Energy Sources and Its Storage in Salt Caverns. Energies 2022, 15, 1415. [Google Scholar] [CrossRef]
  111. de Jongh, P.E.; Adelhelm, P. Nanosizing and nanoconfinement: New strategies towards meeting hydrogen storage goals. ChemSusChem 2010, 3, 1332–1348. [Google Scholar] [CrossRef]
  112. Mazloomi, K.; Gomes, C. Hydrogen as an energy carrier: Prospects and challenges. Renew. Sustain. Energy Rev. 2012, 16, 3024–3033. [Google Scholar] [CrossRef]
  113. Schitea, D.; Deveci, M.; Iordache, M.; Bilgili, K.; Akyurt, I.Z.; Iordache, I. Hydrogen mobility roll-up site selection using intuitionistic fuzzy sets based WASPAS, COPRAS and EDAS. Int. J. Hydrogen Energy 2019, 44, 8585–8600. [Google Scholar] [CrossRef]
  114. Hambourger, M.; Moore, T.A. Nailing down nickel for electrocatalysis. Science 2009, 326, 1355–1356. [Google Scholar] [CrossRef]
  115. Khafidz, N.Z.A.K.; Yaakob, Z.; Lim, K.L.; Timmiati, S.N. The kinetics of lightweight solid-state hydrogen storage materials: A review. Int. J. Hydrogen Energy 2016, 41, 13131–13151. [Google Scholar] [CrossRef]
  116. Jiang, W.; Wang, H.; Zhu, M. AlH3 as a hydrogen storage material: Recent advances, prospects and challenges. Rare Met. 2021, 40, 3337–3356. [Google Scholar] [CrossRef]
  117. Ley, M.B.; Jepsen, L.H.; Lee, Y.-S.; Cho, Y.W.; Von Colbe, J.M.B.; Dornheim, M.; Rokni, M.; Jensen, J.O.; Sloth, M.; Filinchuk, Y. Complex hydrides for hydrogen storage–new perspectives. Mater. Today 2014, 17, 122–128. [Google Scholar] [CrossRef]
  118. Bullock, R.M. Metal-hydrogen bond cleavage reactions of transition metal hydrides: Hydrogen atom, hydride, and proton transfer reactions. Comments Inorg. Chem. 1991, 12, 1–33. [Google Scholar] [CrossRef]
  119. Rönnebro, E.C.; Majzoub, E.H. Recent advances in metal hydrides for clean energy applications. MRS Bull. 2013, 38, 452–458. [Google Scholar] [CrossRef]
  120. Crivello, J.-C.; Denys, R.; Dornheim, M.; Felderhoff, M.; Grant, D.; Huot, J.; Jensen, T.; De Jongh, P.; Latroche, M.; Walker, G. Mg-based compounds for hydrogen and energy storage. Appl. Phys. A 2016, 122, 85. [Google Scholar] [CrossRef]
  121. He, T.; Pachfule, P.; Wu, H.; Xu, Q.; Chen, P. Hydrogen carriers. Nat. Rev. Mater. 2016, 1, 16059. [Google Scholar] [CrossRef]
  122. Mustafa, A.; Lougou, B.G.; Shuai, Y.; Wang, Z.; Tan, H. Current technology development for CO2 utilization into solar fuels and chemicals: A review. J. Energy Chem. 2020, 49, 96–123. [Google Scholar] [CrossRef]
  123. Crabtree, R.H. Hydrogen storage in liquid organic heterocycles. Energy Environ. Sci. 2008, 1, 134–138. [Google Scholar] [CrossRef]
  124. Iguchi, M.; Himeda, Y.; Manaka, Y.; Matsuoka, K.; Kawanami, H. Simple Continuous High-Pressure Hydrogen Production and Separation System from Formic Acid under Mild Temperatures. ChemCatChem 2016, 8, 886–890. [Google Scholar] [CrossRef]
  125. Rahmaninasab, M.A.; Raygan, S.; Abdizadeh, H.; Pourabdoli, M.; Mirghaderi, S.H. Properties of activated MgH2+ mischmetal nanostructured composite produced by ball-milling. Mater. Renew. Sustain. Energy 2018, 7, 15. [Google Scholar] [CrossRef]
  126. Erickson, K.J.; Gibb, A.L.; Sinitskii, A.; Rousseas, M.; Alem, N.; Tour, J.M.; Zettl, A.K. Longitudinal splitting of boron nitride nanotubes for the facile synthesis of high quality boron nitride nanoribbons. Nano Lett. 2011, 11, 3221–3226. [Google Scholar] [CrossRef]
  127. Li, M.; Li, Y.; Zhou, Z.; Shen, P.; Chen, Z. Ca-coated boron fullerenes and nanotubes as superior hydrogen storage materials. Nano Lett. 2009, 9, 1944–1948. [Google Scholar] [CrossRef]
  128. Öberg, S.; Odenberger, M.; Johnsson, F. Exploring the competitiveness of hydrogen-fueled gas turbines in future energy systems. Int. J. Hydrogen Energy 2022, 47, 624–644. [Google Scholar] [CrossRef]
  129. Katja Purr, U.S.; Werner, K.; Nissler, D.; Will, M.; Knoche, G.; Volkens, A. Germany in 2050—A Greenhouse Gas-Neutral Country; Umweltbundesamt: Berlin, Germany, 2014. [Google Scholar]
  130. Hydrogen Gas Turbine Offers Promise of Clean Electricity. Available online: https://0-www-nature-com.brum.beds.ac.uk/articles/d42473-022-00211-0 (accessed on 21 November 2022).
  131. Germany launches world’s first hydrogen-powered train. The Guardian, 17 September 2018.
  132. IEA. World Energy Model Documentation; IEA: Paris, France, 2021. [Google Scholar]
  133. Pörtner, H.O.; Roberts, D.C.; Admas, H.; Adler, C.; Aldunce, P.; Ali, E.; Begum, R.A.; Betts, R.; Kerr, R.B.; Biesbroek, R.; et al. Summary for Policymakers. In Climate Change 2022: Impacts, Adaptation and Vulnerability; Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Pörtner, D.C.R., Poloczanska, E.S., Mintenbeck, K., Tignor, M., Alegría, A., Craig, M., Langsdorf, S., Löschke, S., Möller, V., Okem, A., Eds.; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2022; pp. 3–33. [Google Scholar]
Figure 1. Comparison of specific energy (energy per mass or gravimetric density) and energy density (energy per volume or volumetric density) for several fuels based on lower heating values. After [14].
Figure 1. Comparison of specific energy (energy per mass or gravimetric density) and energy density (energy per volume or volumetric density) for several fuels based on lower heating values. After [14].
Energies 16 01484 g001
Figure 2. Integrated RE H2 system and economy.
Figure 2. Integrated RE H2 system and economy.
Energies 16 01484 g002
Figure 3. The spectrum of H2 production processes.
Figure 3. The spectrum of H2 production processes.
Energies 16 01484 g003
Figure 4. (a) H2 generation from alkaline electrolyzer. Reproduced with permission from [58], John Wiley and Sons, 2022; and (b) schematics of PEM water electrolyzer (Solid polymer electrolyte membrane is the same as proton exchange membrane).
Figure 4. (a) H2 generation from alkaline electrolyzer. Reproduced with permission from [58], John Wiley and Sons, 2022; and (b) schematics of PEM water electrolyzer (Solid polymer electrolyte membrane is the same as proton exchange membrane).
Energies 16 01484 g004
Figure 5. Overview of H2 storage technologies.
Figure 5. Overview of H2 storage technologies.
Energies 16 01484 g005
Table 1. Summary of H2 production technologies.
Table 1. Summary of H2 production technologies.
H2 Production TechnologySourcesEfficiencyAdvantagesCarbon Emission Status
Steam methane reforming (SMR)Natural gas, Hydrocarbons70–85%Cheap, high efficiency, established technologyCO2 emission
Coal gasificationHydrocarbons60–70%Abundance of source material, established technologyCO2 emission
Biomass gasificationBiomass35–50%Renewable technologyCO2 emission
PyrolysisNatural gas, Hydrocarbons78–90%Low carbon emission, efficientSolid C byproduct
Nuclear produced H2Steam-Low carbon emission, no fossil-fuel requiredNo CO2 emission
Electrolysis: AKLWater70%Safe, long lifetime, cheaperNo CO2 emission
Electrolysis: PEMWater60%Low maintenance cost, small, recycling of electrolyte is not requiredNo CO2 emission
Electrolysis: SOSESteam-Flexible application, scalability, pure H2, reverse mode operationNo CO2 emission
Photocatalytic and photoelectrochemical Water and sunlight~5%Electricity independent, self-sufficient No CO2 emission
Table 2. Comparison among different H2 storage technologies.
Table 2. Comparison among different H2 storage technologies.
H2 Storage TechnologiesTemperaturePressureStorage CapacityApplicationsMeritsDemeritsRef.
Pressurized H2 tanksLowHigh0.99H atoms/cm3Short term supply and storage for commercial useEstablished H2 storage facilityHeavy metal tanks; storage density not great[58,109]
Underground salt cavernsLowHigh10 kg/m3Long term storage and emergency backupHigh density energy storageNot suitable for mobile use or transportation[62,110]
Liquid H2CryogenicExtremely high71 g/LTransportation,
automotive, air, and space
Large amount transport, lighter, compact, saferEnergy loss for cooling[111,112,113]
Metal hydridesModerateModerate (3–30 atm)6.5H atoms/cm3Onboard applicationsSafe, high storage density, volume efficient, reversibilityHigh weight and low storage capacity at low temperature[114,115,117,118,119,120]
Chemical carriersNear ambientNear ambient (~10 atm)(99–109) g/LEasy Transportation and suitable for FCV useRelatively dense storage capacity, no need for extra compression systemBy products (N2, CO2), need CCS unit (for one way carrier).[121,122,123,124]
Nano absorbersAmbientAmbient7.6–8.2 wt.%Compatible for weight and size restricted onboard systemsBulk absorption/desorption, good storage capacity, low diffusion activation and binding energyNot suitable for mass storage[62,126,127]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Noyan, O.F.; Hasan, M.M.; Pala, N. A Global Review of the Hydrogen Energy Eco-System. Energies 2023, 16, 1484. https://0-doi-org.brum.beds.ac.uk/10.3390/en16031484

AMA Style

Noyan OF, Hasan MM, Pala N. A Global Review of the Hydrogen Energy Eco-System. Energies. 2023; 16(3):1484. https://0-doi-org.brum.beds.ac.uk/10.3390/en16031484

Chicago/Turabian Style

Noyan, Omer Faruk, Muhammad Mahmudul Hasan, and Nezih Pala. 2023. "A Global Review of the Hydrogen Energy Eco-System" Energies 16, no. 3: 1484. https://0-doi-org.brum.beds.ac.uk/10.3390/en16031484

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop