Hydrogen economy

The hydrogen economy is the use of hydrogen as a low carbon fuel, particularly for heat;[1] but also for hydrogen vehicles,[2] seasonal energy storage and long distance transport of energy.[3]

The hydrogen economy is developing as a small part of the low-carbon economy.[4] In order to phase out fossil fuels and limit global warming, hydrogen is starting to be used as its combustion only releases clean water, and no CO
2
to the atmosphere.[5] As of 2019, however, hydrogen is mainly used as an industrial feedstock, primarily for the production of ammonia, methanol and petroleum refining.

Hydrogen gas does not occur naturally in convenient reservoirs. As of 2019 almost all the world's 70 million tons of hydrogen consumed yearly in industrial processing[6] is produced by steam methane reforming (SMR).[7] Small amounts of hydrogen are produced by the dedicated production of hydrogen from water. As of 2019 there is not enough cheap clean electricity (renewable and nuclear) for this hydrogen to become a significant part of the low-carbon economy, and carbon dioxide is a by-product of the SMR process,[8] but it can be captured and stored.

Rationale

Elements of the hydrogen economy

In the current hydrocarbon economy, heating is fueled primarily by natural gas and transportation by petroleum. Burning of hydrocarbon fuels emits carbon dioxide and other pollutants. The demand for energy is increasing, particularly in China, India, and other developing countries. Hydrogen can be an environmentally cleaner source of energy to end-users, without release of pollutants such as particulates or carbon dioxide.[9]

Hydrogen has a high energy density by weight but has a low energy density by volume. Even when highly compressed, stored in solids, or liquified, the energy density by volume is only 1/4 that of gasoline, although the energy density by weight is approximately three times that of gasoline or natural gas. Hydrogen can help to decarbonize long-haul transport, chemicals, and iron and steel[3] and has the potential to transport renewable energy long distance and store it long term, for example from wind power or solar electricity.[10]

History

The term hydrogen economy was coined by John Bockris during a talk he gave in 1970 at General Motors (GM) Technical Center.[11] The concept was proposed earlier by geneticist J.B.S. Haldane.[12]

A hydrogen economy was proposed by the University of Michigan to solve some of the negative effects of using hydrocarbon fuels where the carbon is released to the atmosphere (as carbon dioxide, carbon monoxide, unburnt hydrocarbons, etc.). Modern interest in the hydrogen economy can generally be traced to a 1970 technical report by Lawrence W. Jones of the University of Michigan.[13]

A spike in attention for the concept during the 2000s was repeatedly described as hype by some critics and proponents of alternative technologies.[14][15][16] İnterest in the energy carrier resurged in the 2010s, notably by the forming of the Hydrogen Council in 2017. Several manufacturers released hydrogen fuel cell cars commercially, with manufacturers such as Toyota and industry groups in China planning to increase numbers of the cars into the hundreds of thousands over the next decade.[17][18]

Current hydrogen market

Timeline

Hydrogen production is a large and growing industry: with as of 2019 about 70 million tonnes of dedicated production per year, larger than the primary energy supply of Germany.[19]

As of 2019 fertiliser production and oil refining are the main uses.[20] About half is used in the Haber process to produce ammonia (NH3), which is then used directly or indirectly as fertilizer.[21] Because both the world population and the intensive agriculture used to support it are growing, ammonia demand is growing. Ammonia can be used as a safer and easier indirect method of transporting hydrogen. Transported ammonia can be then converted back to hydrogen at the bowser by a membrane technology.[22]

The other half of current hydrogen production is used to convert heavy petroleum sources into lighter fractions suitable for use as fuels. This latter process is known as hydrocracking. Hydrocracking represents an even larger growth area, since rising oil prices encourage oil companies to extract poorer source material, such as oil sands and oil shale. The scale economies inherent in large-scale oil refining and fertilizer manufacture make possible on-site production and "captive" use. Smaller quantities of "merchant" hydrogen are manufactured and delivered to end users as well.

As of 2019 almost all hydrogen production is from fossil fuels, and emits 830 million tonnes of carbon dioxide per year.[19] The distribution of production reflects the effects of thermodynamic constraints on economic choices: of the four methods for obtaining hydrogen, partial combustion of natural gas in a NGCC (natural gas combined cycle) power plant offers the most efficient chemical pathway and the greatest off-take of usable heat energy.

The large market and sharply rising prices in fossil fuels have also stimulated great interest in alternate, cheaper means of hydrogen production.[23][24] As of 2002, most hydrogen is produced on site and the cost is approximately $0.70/kg and, if not produced on site, the cost of liquid hydrogen is about $2.20/kg to $3.08/kg.[25]

Production, storage, infrastructure

Today's hydrogen is mainly produced (>90%) from fossil sources.[26]

Methods of production

Molecular hydrogen is not available on Earth in convenient natural reservoirs. Most hydrogen in the lithosphere is bonded to oxygen in water. Manufacturing elemental hydrogen requires the consumption of a hydrogen carrier such as a fossil fuel or water. The former carrier consumes the fossil resource and produces carbon dioxide, but often requires no further energy input beyond the fossil fuel. Decomposing water, the latter carrier, requires electrical or heat input, generated from some primary energy source (fossil fuel, nuclear power or a renewable energy). Hydrogen can also be produced by refining the effluent from geothermal sources in the lithosphere. Hydrogen produced by zero emission energy sources such as electrolysis of water using wind power, solar power, nuclear power, hydro power, wave power or tidal power is referred to as green hydrogen.[27] Hydrogen produced from coal may be referred to as brown hydrogen,[28], and when fossil fuel derived, is generally referred to as grey hydrogen. When derived from natural gas, if the carbon dioxide is captured, it is referred to as blue hydrogen.[29]

Current production methods

Hydrogen is industrially produced from steam reforming, which uses natural gas.[30] The energy content of the produced hydrogen is less than the energy content of the original fuel, some of it being lost as excess heat during production. Steam reforming emits carbon dioxide.

A small part (2% in 2019[31]) is produced by electrolysis using electricity and water, consuming approximately 50 to 55 kilowatt-hours of electricity per kilogram of hydrogen produced.[32]

Electrolysis of water

Hydrogen can be made via high pressure electrolysis, low pressure electrolysis of water, or a range of other emerging electrochemical processes such as high temperature electrolysis or carbon assisted electrolysis.[33] However, current best processes for water electrolysis have an effective electrical efficiency of 70-80%,[34][35][36] so that producing 1 kg of hydrogen (which has a specific energy of 143 MJ/kg or about 40 kWh/kg) requires 50–55 kWh of electricity.

In parts of the world, steam methane reforming is between $1–3/kg on average. This makes production of hydrogen via electrolysis cost competitive in many regions already, as outlined by Nel Hydrogen[37] and others, including an article by the IEA[38] examining the conditions which could lead to a competitive advantage for electrolysis.

Kværner process

The Kværner process or Kvaerner carbon black & hydrogen process (CB&H)[26] is a method, developed in the 1980s by a Norwegian company of the same name, for the production of hydrogen from hydrocarbons (CnHm), such as methane, natural gas and biogas. Of the available energy of the feed, approximately 48% is contained in the hydrogen, 40% is contained in activated carbon and 10% in superheated steam.[39]

Experimental production methods

Biological production

Fermentative hydrogen production is the fermentative conversion of organic substrate to biohydrogen manifested by a diverse group of bacteria using multi enzyme systems involving three steps similar to anaerobic conversion. Dark fermentation reactions do not require light energy, so they are capable of constantly producing hydrogen from organic compounds throughout the day and night. Photofermentation differs from dark fermentation because it only proceeds in the presence of light. For example, photo-fermentation with Rhodobacter sphaeroides SH2C can be employed to convert small molecular fatty acids into hydrogen.[40] Electrohydrogenesis is used in microbial fuel cells where hydrogen is produced from organic matter (e.g. from sewage, or solid matter[41]) while 0.2 - 0.8 V is applied.

Biological hydrogen can be produced in an algae bioreactor. In the late 1990s it was discovered that if the algae is deprived of sulfur it will switch from the production of oxygen, i.e. normal photosynthesis, to the production of hydrogen.[42]

Biological hydrogen can be produced in bioreactors that use feedstocks other than algae, the most common feedstock being waste streams. The process involves bacteria feeding on hydrocarbons and excreting hydrogen and CO2. The CO2 can be sequestered successfully by several methods, leaving hydrogen gas. In 2006-2007, NanoLogix first demonstrated a prototype hydrogen bioreactor using waste as a feedstock at Welch's grape juice factory in North East, Pennsylvania (U.S.).[43]

Biocatalysed electrolysis

Besides regular electrolysis, electrolysis using microbes is another possibility. With biocatalysed electrolysis, hydrogen is generated after running through the microbial fuel cell and a variety of aquatic plants can be used. These include reed sweetgrass, cordgrass, rice, tomatoes, lupines, and algae[44]

High-pressure electrolysis

High pressure electrolysis is the electrolysis of water by decomposition of water (H2O) into oxygen (O2) and hydrogen gas (H2) by means of an electric current being passed through the water. The difference with a standard electrolyzer is the compressed hydrogen output around 120-200 bar (1740-2900 psi, 12–20 MPa).[45] By pressurising the hydrogen in the electrolyser, through a process known as chemical compression, the need for an external hydrogen compressor is eliminated,[46] the average energy consumption for internal compression is around 3%.[47] European largest (1 400 000 kg/a, High-pressure Electrolysis of water, alkaline technology) hydrogen production plant is operating at Kokkola, Finland.[48]

High-temperature electrolysis

Hydrogen can be generated from energy supplied in the form of heat and electricity through high-temperature electrolysis (HTE). Because some of the energy in HTE is supplied in the form of heat, less of the energy must be converted twice (from heat to electricity, and then to chemical form), and so potentially far less energy is required per kilogram of hydrogen produced.

While nuclear-generated electricity could be used for electrolysis, nuclear heat can be directly applied to split hydrogen from water. High temperature (950–1000 °C) gas cooled nuclear reactors have the potential to split hydrogen from water by thermochemical means using nuclear heat. Research into high-temperature nuclear reactors may eventually lead to a hydrogen supply that is cost-competitive with natural gas steam reforming. General Atomics predicts that hydrogen produced in a High Temperature Gas Cooled Reactor (HTGR) would cost $1.53/kg. In 2003, steam reforming of natural gas yielded hydrogen at $1.40/kg. In 2005 natural gas prices, hydrogen costs $2.70/kg.

High-temperature electrolysis has been demonstrated in a laboratory, at 108 MJ (thermal) per kilogram of hydrogen produced,[49] but not at a commercial scale. In addition, this is lower-quality "commercial" grade Hydrogen, unsuitable for use in fuel cells.[50]

Photoelectrochemical water splitting

Using electricity produced by photovoltaic systems offers the cleanest way to produce hydrogen. Water is broken into hydrogen and oxygen by electrolysis—a photoelectrochemical cell (PEC) process which is also named artificial photosynthesis.[51] William Ayers at Energy Conversion Devices demonstrated and patented the first multijunction high efficiency photoelectrochemical system for direct splitting of water in 1983.[52] This group demonstrated direct water splitting now referred to as an "artificial leaf" or "wireless solar water splitting" with a low cost thin film amorphous silicon multijunction sheet immersed directly in water. Hydrogen evolved on the front amorphous silicon surface decorated with various catalysts while oxygen evolved off the back metal substrate. A Nafion membrane above the multijunction cell provided a path for ion transport. Their patent also lists a variety of other semiconductor multijunction materials for the direct water splitting in addition to amorphous silicon and silicon germanium alloys. Research continues towards developing high-efficiency multi-junction cell technology at universities and the photovoltaic industry. If this process is assisted by photocatalysts suspended directly in water instead of using photovoltaic and an electrolytic system, the reaction is in just one step, which can improve efficiency.[53][54]

Photoelectrocatalytic production

A method studied by Thomas Nann and his team at the University of East Anglia consists of a gold electrode covered in layers of indium phosphide (InP) nanoparticles. They introduced an iron-sulfur complex into the layered arrangement, which when submerged in water and irradiated with light under a small electric current, produced hydrogen with an efficiency of 60%.[55]

In 2015, it was reported that Panasonic Corp. has developed a photocatalyst based on niobium nitride that can absorb 57% of sunlight to support the decomposition of water to produce hydrogen gas.[56] The company plans to achieve commercial application "as early as possible", not before 2020.

Concentrating solar thermal

Very high temperatures are required to dissociate water into hydrogen and oxygen. A catalyst is required to make the process operate at feasible temperatures. Heating the water can be achieved through the use of concentrating solar power. Hydrosol-2 is a 100-kilowatt pilot plant at the Plataforma Solar de Almería in Spain which uses sunlight to obtain the required 800 to 1,200 °C to heat water. Hydrosol II has been in operation since 2008. The design of this 100-kilowatt pilot plant is based on a modular concept. As a result, it may be possible that this technology could be readily scaled up to the megawatt range by multiplying the available reactor units and by connecting the plant to heliostat fields (fields of sun-tracking mirrors) of a suitable size.[57]

Thermochemical production

There are more than 352[58] thermochemical cycles which can be used for water splitting,[59] around a dozen of these cycles such as the iron oxide cycle, cerium(IV) oxide-cerium(III) oxide cycle, zinc zinc-oxide cycle, sulfur-iodine cycle, copper-chlorine cycle and hybrid sulfur cycle are under research and in testing phase to produce hydrogen and oxygen from water and heat without using electricity.[60] These processes can be more efficient than high-temperature electrolysis, typical in the range from 35% - 49% LHV efficiency. Thermochemical production of hydrogen using chemical energy from coal or natural gas is generally not considered, because the direct chemical path is more efficient.

None of the thermochemical hydrogen production processes have been demonstrated at production levels, although several have been demonstrated in laboratories.

Hydrogen as a byproduct of other chemical processes

The industrial production of chlorine and caustic soda by electrolysis generates a sizable amount of Hydrogen as a byproduct. In the port of Antwerp a 1MW demonstration fuel cell power plant is powered by such byproduct. This unit has been operational since late 2011.[61] The excess hydrogen is often managed with a hydrogen pinch analysis.

Storage

Although molecular hydrogen has very high energy density on a mass basis, partly because of its low molecular weight, as a gas at ambient conditions it has very low energy density by volume. If it is to be used as fuel stored on board the vehicle, pure hydrogen gas must be stored in an energy-dense form to provide sufficient driving range.

Pressurized hydrogen gas

Increasing gas pressure improves the energy density by volume making for smaller container tanks. Tanks made of carbon and glass fibres reinforcing plastic as fitted in Toyota Marai and Kenworth trucks are required to meet safety standards. Few materials are suitable for tanks as hydrogen being a small molecule tends to diffuse through many liner materials and hydrogen embrittlement causes weakening in some types of metal containers. The most common on board hydrogen storage in today's 2020 vehicles is hydrogen at pressure 700bar = 70MPa.

Liquid hydrogen

Alternatively, higher volumetric energy density liquid hydrogen or slush hydrogen may be used. However, liquid hydrogen is cryogenic and boils at 20.268 K (–252.882 °C or –423.188 °F). Cryogenic storage cuts weight but requires large liquification energies. The liquefaction process, involving pressurizing and cooling steps, is energy intensive.[62] The liquefied hydrogen has lower energy density by volume than gasoline by approximately a factor of four, because of the low density of liquid hydrogen — there is actually more hydrogen in a litre of gasoline (116 grams) than there is in a litre of pure liquid hydrogen (71 grams). Liquid hydrogen storage tanks must also be well insulated to minimize boil off.

Japan have a liquid hydrogen (LH2) storage facility at a terminal in Kobe, and are expected to receive the first shipment of liquid hydrogen via LH2 carrier in 2020.[63] Hydrogen is liquified by reducing its temperature to -253 °C, similar to liquified natural gas (LNG) which is stored at -162 °C. A potential efficiency loss of 12.79% can be achieved, or 4.26kWh/kg out of 33.3kWh/kg.[64]

Storage as hydride

Distinct from storing molecular hydrogen, hydrogen can be stored as a chemical hydride or in some other hydrogen-containing compound. Hydrogen gas is reacted with some other materials to produce the hydrogen storage material, which can be transported relatively easily. At the point of use the hydrogen storage material can be made to decompose, yielding hydrogen gas. As well as the mass and volume density problems associated with molecular hydrogen storage, current barriers to practical storage schemes stem from the high pressure and temperature conditions needed for hydride formation and hydrogen release. For many potential systems hydriding and dehydriding kinetics and heat management are also issues that need to be overcome. A French company McPhy Energy is developing the first industrial product, based on Magnesium Hydrate, already sold to some major clients such as Iwatani and ENEL. Emergent hydride hydrogen storage technologies have achieved a compressed volume of less than 1/500.

Adsorption

A third approach is to adsorb molecular hydrogen on the surface of a solid storage material. Unlike in the hydrides mentioned above, the hydrogen does not dissociate/recombine upon charging/discharging the storage system, and hence does not suffer from the kinetic limitations of many hydride storage systems. Hydrogen densities similar to liquefied hydrogen can be achieved with appropriate adsorbent materials. Some suggested adsorbents include activated carbon, nanostructured carbons (including CNTs), MOFs, and hydrogen clathrate hydrate.

Underground hydrogen storage

'Available storage technologies, their capacity and discharge time.' COMMISSION STAFF WORKING DOCUMENT Energy storage – the role of electricity

Underground hydrogen storage is the practice of hydrogen storage in caverns, salt domes and depleted oil and gas fields. Large quantities of gaseous hydrogen have been stored in caverns by ICI for many years without any difficulties.[65] The storage of large quantities of liquid hydrogen underground can function as grid energy storage. The round-trip efficiency is approximately 40% (vs. 75-80% for pumped-hydro (PHES)), and the cost is slightly higher than pumped hydro.[66] Another study referenced by a European staff working paper found that for large scale storage, the cheapest option is hydrogen at €140/MWh for 2,000 hours of storage using an electrolyser, salt cavern storage and combined-cycle power plant.[67] The European project Hyunder[68] indicated in 2013 that for the storage of wind and solar energy an additional 85 caverns are required as it cannot be covered by PHES and CAES systems.[69] A German case study on storage of hydrogen in salt caverns found that if the German power surplus (7% of total variable renewable generation by 2025 and 20% by 2050) would be converted to hydrogen and stored underground, these quantities would require some 15 caverns of 500,000 cubic metres each by 2025 and some 60 caverns by 2050 – corresponding to approximately one third of the number of gas caverns currently operated in Germany.[70] In the US, Sandia Labs are conducting research into the storage of hydrogen in depleted oil and gas fields, which could easily absorb large amounts of renewably produced hydrogen as there are some 2.7 million depleted wells in existence.[71]

Power to gas

Power to gas is a technology which converts electrical power to a gas fuel. There are 2 methods, the first is to use the electricity for water splitting and inject the resulting hydrogen into the natural gas grid. The second (less efficient) method is used to convert carbon dioxide and water to methane, (see natural gas) using electrolysis and the Sabatier reaction. The excess power or off peak power generated by wind generators or solar arrays is then used for load balancing in the energy grid. Using the existing natural gas system for hydrogen Fuel cell maker Hydrogenics and natural gas distributor Enbridge have teamed up to develop such a power to gas system in Canada.[72]

Pipeline storage

A natural gas network may be used for the storage of hydrogen. Before switching to natural gas, the UK and German gas networks were operated using towngas, which for the most part consisted of hydrogen. The storage capacity of the German natural gas network is more than 200,000 GWh which is enough for several months of energy requirement. By comparison, the capacity of all German pumped storage power plants amounts to only about 40 GW·h. Similarly UK pumped storage is far less than the gas network. The transport of energy through a gas network is done with much less loss (<0.1%) than in a power network (8%). The use of the existing natural gas pipelines for hydrogen was studied by NaturalHy.[73] Ad van Wijk, a professor at Future Energy Systems TU Delft, also discusses the possibility of producing electricity in areas or countries with much sunlight (Sahara, Chile, Mexico, Namibia, Australia, New Zealand, ...) and transporting it (via ship, pipeline, ...) to the Netherlands. This being economically seen, still cheaper than producing it locally in the Netherlands. He also mentions that the energy transport capacity of gas lines are far higher than that of electricity lines coming into private houses (in the Netherlands) -30 kW vs 3 kW-.[74][75]

Infrastructure

Praxair Hydrogen Plant

The hydrogen infrastructure would consist mainly of industrial hydrogen pipeline transport and hydrogen-equipped filling stations like those found on a hydrogen highway. Hydrogen stations which were not situated near a hydrogen pipeline would get supply via hydrogen tanks, compressed hydrogen tube trailers, liquid hydrogen trailers, liquid hydrogen tank trucks or dedicated onsite production.

Because of hydrogen embrittlement of steel, and corrosion[76][77] natural gas pipes require internal coatings or replacement in order to convey hydrogen. Techniques are well-known; over 700 miles of hydrogen pipeline currently exist in the United States. Although expensive, pipelines are the cheapest way to move hydrogen. Hydrogen gas piping is routine in large oil-refineries, because hydrogen is used to hydrocrack fuels from crude oil.

Hydrogen piping can in theory be avoided in distributed systems of hydrogen production, where hydrogen is routinely made on site using medium or small-sized generators which would produce enough hydrogen for personal use or perhaps a neighborhood. In the end, a combination of options for hydrogen gas distribution may succeed.

The IEA recommends existing industrial ports be used for production and existing natural gas pipelines for transport: also international co-operation and shipping.[78]

South Korea and Japan,[79] which as of 2019 lack international electrical interconnectors, are investing in the hydrogen economy.[80] In March 2020, a production facility was opened in Namie, Fukushima Prefecture, claimed to be the world's largest.[81]

A key tradeoff: centralized vs. distributed production

In a future full hydrogen economy, primary energy sources and feedstock would be used to produce hydrogen gas as stored energy for use in various sectors of the economy. Producing hydrogen from primary energy sources other than coal, oil, and natural gas, would result in lower production of the greenhouse gases characteristic of the combustion of these fossil energy resources.

One key feature of a hydrogen economy would be that in mobile applications (primarily vehicular transport) energy generation and use could be decoupled. The primary energy source would need no longer travel with the vehicle, as it currently does with hydrocarbon fuels. Instead of tailpipes creating dispersed emissions, the energy (and pollution) could be generated from point sources such as large-scale, centralized facilities with improved efficiency. This would allow the possibility of technologies such as carbon sequestration, which are otherwise impossible for mobile applications. Alternatively, distributed energy generation schemes (such as small scale renewable energy sources) could be used, possibly associated with hydrogen stations.

Aside from the energy generation, hydrogen production could be centralized, distributed or a mixture of both. While generating hydrogen at centralized primary energy plants promises higher hydrogen production efficiency, difficulties in high-volume, long range hydrogen transportation (due to factors such as hydrogen damage and the ease of hydrogen diffusion through solid materials) makes electrical energy distribution attractive within a hydrogen economy. In such a scenario, small regional plants or even local filling stations could generate hydrogen using energy provided through the electrical distribution grid. While hydrogen generation efficiency is likely to be lower than for centralized hydrogen generation, losses in hydrogen transport could make such a scheme more efficient in terms of the primary energy used per kilogram of hydrogen delivered to the end user.

The proper balance between hydrogen distribution and long-distance electrical distribution is one of the primary questions that arises about the hydrogen economy.

Again the dilemmas of production sources and transportation of hydrogen can now be overcome using on site (home, business, or fuel station) generation of hydrogen from off grid renewable sources..

Distributed electrolysis

Distributed electrolysis would bypass the problems of distributing hydrogen by distributing electricity instead. It would use existing electrical networks to transport electricity to small, on-site electrolysers located at filling stations. However, accounting for the energy used to produce the electricity and transmission losses would reduce the overall efficiency.

Uses

For heating and cooking instead of natural gas

Hydrogen can replace some or all of the natural gas in gas grids.[82] As of 2020 the maximum in a grid is 20%.[83]

Fuel cells as alternative to internal combustion and electric batteries

One of the main offerings of a hydrogen economy is that the fuel can replace the fossil fuel burned in internal combustion engines and turbines as the primary way to convert chemical energy into kinetic or electrical energy, thereby eliminating greenhouse gas emissions and pollution from that engine. Ad van Wijk, a professor at Future Energy Systems TU Delft also mentions that hydrogen is better for larger vehicles - such as trucks, buses and ships - than electric batteries.[84] This because a 1 kg battery, as of 2019, can store 0.1 kWh of energy whereas 1 kg of hydrogen has a usable capacity of 33 kWh.[85]

Although hydrogen can be used in conventional internal combustion engines, fuel cells, being electrochemical, have a theoretical efficiency advantage over heat engines. Fuel cells are more expensive to produce than common internal combustion engines.

Some types of fuel cells work with hydrocarbon fuels,[86] while all can be operated on pure hydrogen. In the event that fuel cells become price-competitive with internal combustion engines and turbines, large gas-fired power plants could adopt this technology.

Hydrogen gas must be distinguished as "technical-grade" (five nines pure, 99.999%), which is suitable for applications such as fuel cells, and "commercial-grade", which has carbon- and sulfur-containing impurities, but which can be produced by the much cheaper steam-reformation process. Fuel cells require high-purity hydrogen because the impurities would quickly degrade the life of the fuel cell stack.

Much of the interest in the hydrogen economy concept is focused on the use of fuel cells to power hydrogen vehicles. Current hydrogen fuel cells suffer from a low power-to-weight ratio.[87] Fuel cells are much more efficient than internal combustion engines, and produce no harmful emissions. If a practical method of hydrogen storage is introduced, and fuel cells become cheaper, they can be economically viable to power hybrid fuel cell/battery vehicles, or purely fuel cell-driven ones. The economic viability of fuel cell powered vehicles will improve as the use of internal combustion engine vehicles becomes more expensive because of charges to cover the costs of their air pollution, through such measures as carbon taxes and low-emission zones.

The combination of the fuel cell and electric motor is 2-3 times more efficient than an internal-combustion engine.[88] Capital costs of fuel cells have reduced significantly over recent years, with a modeled cost of $50/kW cited by the Department of Energy.[89]

Previous technical obstacles have included hydrogen storage issues[90] and the purity requirement of hydrogen used in fuel cells, as with current technology, an operating fuel cell requires the purity of hydrogen to be as high as 99.999%.

Other fuel cell technologies based on the exchange of metal ions (e.g. zinc-air fuel cells) are typically more efficient at energy conversion than hydrogen fuel cells, but the widespread use of any electrical energy → chemical energy → electrical energy systems would necessitate the production of electricity.

Use as a transport fuel and system efficiency

An accounting of the energy utilized during a thermodynamic process, known as an energy balance, can be applied to automotive fuels. With today's technology, the manufacture of hydrogen via steam reforming can be accomplished with a thermal efficiency of 75 to 80 percent. Additional energy will be required to liquefy or compress the hydrogen, and to transport it to the filling station via truck or pipeline. The energy that must be utilized per kilogram to produce, transport and deliver hydrogen (i.e., its well-to-tank energy use) is approximately 50 MJ using technology available in 2004. Subtracting this energy from the enthalpy of one kilogram of hydrogen, which is 141 MJ, and dividing by the enthalpy, yields a thermal energy efficiency of roughly 60%.[91] Gasoline, by comparison, requires less energy input, per gallon, at the refinery, and comparatively little energy is required to transport it and store it owing to its high energy density per gallon at ambient temperatures. Well-to-tank, the supply chain for gasoline is roughly 80% efficient (Wang, 2002). Another grid-based method of supplying hydrogen would be to use electrical to run electrolysers. Roughly 6% of electricity is lost during transmission along power lines, and the process of converting the fossil fuel to electricity in the first place is roughly 33 percent efficient.[92][93] Thus if efficiency is the key determinant it would be unlikely hydrogen vehicles would be fueled by such a method, and indeed viewed this way, electric vehicles would appear to be a better choice. However, as noted above, hydrogen can be produced from a number of feedstocks, in centralized or distributed fashion, and these afford more efficient pathways to produce and distribute the fuel.

A study of the well-to-wheels efficiency of hydrogen vehicles compared to other vehicles in the Norwegian energy system indicates that hydrogen fuel-cell vehicles (FCV) tend to be about a third as efficient as EVs when electrolysis is used, with hydrogen Internal Combustion Engines (ICE) being barely a sixth as efficient. Even in the case where hydrogen fuel cells get their hydrogen from natural gas reformation rather than electrolysis, and EVs get their power from a natural gas power plant, the EVs still come out ahead 35% to 25% (and only 13% for a H2 ICE). This compares to 14% for a gasoline ICE, 27% for a gasoline ICE hybrid, and 17% for a diesel ICE, also on a well-to-wheels basis.[94]

Hydrogen has been called one of the least efficient and most expensive possible replacements for gasoline (petrol) in terms of reducing greenhouse gases; other technologies may be less expensive and more quickly implemented.[95][96] A comprehensive study of hydrogen in transportation applications has found that "there are major hurdles on the path to achieving the vision of the hydrogen economy; the path will not be simple or straightforward".[97] Although Ford Motor Company and French Renault-Nissan cancelled their hydrogen car R&D efforts in 2008 and 2009, respectively,[98][99] they signed a 2009 letter of intent with the other manufacturers and Now GMBH in September 2009 supporting the commercial introduction of FCVs by 2015.[100] A study by The Carbon Trust for the UK Department of Energy and Climate Change suggests that hydrogen technologies have the potential to deliver UK transport with near-zero emissions whilst reducing dependence on imported oil and curtailment of renewable generation. However, the technologies face very difficult challenges, in terms of cost, performance and policy. [101] An Otto-cycle internal-combustion engine running on hydrogen is said to have a maximum efficiency of about 38%, 8% higher than a gasoline internal-combustion engine.[102]

In the short term hydrogen has been proposed as a method of reducing harmful diesel exhaust.[103]

Safety

Hydrogen has one of the widest explosive/ignition mix range with air of all the gases with few exceptions such as acetylene, silane, and ethylene oxide. That means that whatever the mix proportion between air and hydrogen, a hydrogen leak will most likely lead to an explosion, not a mere flame, when a flame or spark ignites the mixture. This makes the use of hydrogen particularly dangerous in enclosed areas such as tunnels or underground parking.[104] Pure hydrogen-oxygen flames burn in the ultraviolet color range and are nearly invisible to the naked eye, so a flame detector is needed to detect if a hydrogen leak is burning. Hydrogen is odorless and leaks cannot be detected by smell.

Hydrogen codes and standards are codes and standards for hydrogen fuel cell vehicles, stationary fuel cell applications and portable fuel cell applications. There are codes and standards for the safe handling and storage of hydrogen, for example the standard for the installation of stationary fuel cell power systems from the National Fire Protection Association.

Codes and standards have repeatedly been identified as a major institutional barrier to deploying hydrogen technologies and developing a hydrogen economy. As of 2019 international standards are needed transport, storage and traceability of environmental impact.[3]

One of the measures on the roadmap is to implement higher safety standards like early leak detection with hydrogen sensors.[105] The Canadian Hydrogen Safety Program concluded that hydrogen fueling is as safe as, or safer than, compressed natural gas (CNG) fueling.[106] The European Commission has funded the first higher educational program in the world in hydrogen safety engineering at the University of Ulster. It is expected that the general public will be able to use hydrogen technologies in everyday life with at least the same level of safety and comfort as with today's fossil fuels.

Costs

H2 production cost ($-gge untaxed) at varying natural gas prices

Although much of an existing natural gas grid could be reused with 100% hydrogen, eliminating natural gas from a large area such as Britain would require huge investment.[1] And switching from natural gas to low-carbon heating is more costly if the carbon costs of natural gas are not reflected in its price.[107]

Power plant capacity that now goes unused at night could be used to produce green hydrogen, but this would not be enough,.[108] therefore blue hydrogen with carbon capture and storage is needed, possibly after autothermal reforming of methane rather than steam methane reforming,[1]

As of 2020 green hydrogen costs between $2.50-6.80 per kilogram and blue hydrogen $1.40-2.40/kg compared with high-carbon grey hydrogen at $1–1.80/kg.[108] Deployment of hydrogen can provide a cost-effective option to displace fossil fuels in applications where emissions reductions would otherwise be impractical and/or expensive.[109] These may include heat for buildings and industry, conversion of natural gas-fired power stations,[110] and fuel for aviation and possibly heavy trucks.[111]

Examples and pilot programs

A Mercedes-Benz O530 Citaro powered by hydrogen fuel cells, in Brno, Czech Republic.

Several domestic U.S. automobile manufactures have committed to develop vehicles using hydrogen. The distribution of hydrogen for the purpose of transportation is currently being tested around the world, particularly in the US (California, Massachusetts), Canada, Japan, the EU (Portugal, Norway, Denmark, Germany), and Iceland, but the cost is very high.

The United States have their own hydrogen policy. A joint venture between NREL and Xcel Energy is combining wind power and hydrogen power in the same way in Colorado.[112] Hydro in Newfoundland and Labrador are converting the current wind-diesel Power System on the remote island of Ramea into a Wind-Hydrogen Hybrid Power Systems facility.[113] A similar pilot project on Stuart Island uses solar power, instead of wind power, to generate electricity. When excess electricity is available after the batteries are fully charged, hydrogen is generated by electrolysis and stored for later production of electricity by fuel cell.[114] The US also have a large natural gas pipeline system already in place.[115]

Countries in the EU which have a relatively large natural gas pipeline system already in place include Belgium, Germany, France, and the Netherlands.[115]

The UK started a fuel cell pilot program in January 2004, the program ran two Fuel cell buses on route 25 in London until December 2005, and switched to route RV1 until January 2007.[116] The Hydrogen Expedition is currently working to create a hydrogen fuel cell-powered ship and using it to circumnavigate the globe, as a way to demonstrate the capability of hydrogen fuel cells.[117]

Western Australia's Department of Planning and Infrastructure operated three Daimler Chrysler Citaro fuel cell buses as part of its Sustainable Transport Energy for Perth Fuel Cells Bus Trial in Perth.[118] The buses were operated by Path Transit on regular Transperth public bus routes. The trial began in September 2004 and concluded in September 2007. The buses' fuel cells used a proton exchange membrane system and were supplied with raw hydrogen from a BP refinery in Kwinana, south of Perth. The hydrogen was a byproduct of the refinery's industrial process. The buses were refueled at a station in the northern Perth suburb of Malaga.

Iceland has committed to becoming the world's first hydrogen economy by the year 2050.[119] Iceland is in a unique position. Presently, it imports all the petroleum products necessary to power its automobiles and fishing fleet. Iceland has large geothermal resources, so much that the local price of electricity actually is lower than the price of the hydrocarbons that could be used to produce that electricity.

Iceland already converts its surplus electricity into exportable goods and hydrocarbon replacements. In 2002, it produced 2,000 tons of hydrogen gas by electrolysis, primarily for the production of ammonia (NH3) for fertilizer. Ammonia is produced, transported, and used throughout the world, and 90% of the cost of ammonia is the cost of the energy to produce it.

Neither industry directly replaces hydrocarbons. Reykjavík, Iceland, had a small pilot fleet of city buses running on compressed hydrogen,[120] and research on powering the nation's fishing fleet with hydrogen is under way (for example by companies as Icelandic New Energy). For more practical purposes, Iceland might process imported oil with hydrogen to extend it, rather than to replace it altogether.

The Reykjavík buses are part of a larger program, HyFLEET:CUTE,[121] operating hydrogen fueled buses in eight European cities. HyFLEET:CUTE buses were also operated in Beijing, China and Perth, Australia (see below). A pilot project demonstrating a hydrogen economy is operational on the Norwegian island of Utsira. The installation combines wind power and hydrogen power. In periods when there is surplus wind energy, the excess power is used for generating hydrogen by electrolysis. The hydrogen is stored, and is available for power generation in periods when there is little wind.

India is said to adopt hydrogen and H-CNG, due to several reasons, amongst which the fact that a national rollout of natural gas networks is already taking place and natural gas is already a major vehicle fuel. In addition, India suffers from extreme air pollution in urban areas.[122][123] Currently however, hydrogen energy is just at the Research, Development and Demonstration (RD&D) stage.[124][125] As a result, the number of hydrogen stations may still be low,[126] although much more are expected to be introduced soon.[127][128][129]

The Turkish Ministry of Energy and Natural Resources and the United Nations Industrial Development Organization have signed a $40 million trust fund agreement in 2003 for the creation of the International Centre for Hydrogen Energy Technologies (UNIDO-ICHET) in Istanbul, which started operation in 2004.[130] A hydrogen forklift, a hydrogen cart and a mobile house powered by renewable energies are being demonstrated in UNIDO-ICHET's premises. An uninterruptible power supply system has been working since April 2009 in the headquarters of Istanbul Sea Buses company.

Another indicator of the presence of large natural gas infrastructures already in place in countries and in use by citizens is the number of natural gas vehicles present in the country. The countries with the largest amount of natural gas vehicles are (in order of magnitude):[131] Iran, China, Pakistan, Argentina, India, Brasil, Italy, Colombia, Thailand, Uzbekistan, Bolivia, Armenia, Bangladesh, Egypt, Peru, Ukraine, United States. Natural gas vehicles can also be converted to run on hydrogen.

Some hospitals have installed combined electrolyser-storage-fuel cell units for local emergency power. These are advantageous for emergency use because of their low maintenance requirement and ease of location compared to internal combustion driven generators.

Also, in some private homes, fuel cell micro-CHP plants can be found, which can operate on hydrogen, or other fuels as natural gas or LPG.[132][133] When running on natural gas, it relies on steam reforming of natural gas to convert the natural gas to hydrogen prior to use in the fuel cell. This hence still emits CO2 (see reaction) but (temporarily) running on this can be a good solution until the point where the hydrogen is starting to be become distributed through the (natural gas) piping system.

Partial hydrogen economy

Hydrogen is simply a method to store and transmit energy. Energy development of various alternative energy transmission and storage scenarios which begin with hydrogen production, but do not use it for all parts of the store and transmission infrastructure, may be more economic, in both near and far term. These include:

Ammonia economy

An alternative to gaseous hydrogen as an energy carrier is to bond it with nitrogen from the air to produce ammonia, which can be easily liquefied, transported, and used (directly or indirectly) as a clean and renewable fuel.[134][135] For example, researchers at CSIRO in Australia in 2018 fuelled a Toyota Mirai and Hyundai Nexo with hydrogen separated from ammonia using a membrane technology.[22]

Hybrid heat pumps

Hybrid heat pumps (not to be confused with air water hybrids) also include a boiler which could run on methane or hydrogen, and could be a pathway to full decarbonisation of residential heating as the boiler would be used to top up the heating when the weather was very cold.[136]

Bio-SNG

As of 2019 although technically possible production of syngas from hydrogen and carbon-dioxide from bio-energy with carbon capture and storage (BECCS) via the Sabatier reaction is limited by the amount of sustainable bioenergy available:[137] therefore any bio-SNG made may be reserved for production of aviation biofuel.[138]

See also

References

  1. "Transitioning to hydrogen: Assessing the engineering risks and uncertainties". theiet.org. Retrieved 2020-04-11.
  2. A portfolio of power-trains for Europe: a fact-based analysis
  3. IEA H2 2019, p. 13
  4. Deign, Jason (2019-10-14). "10 Countries Moving Toward a Green Hydrogen Economy". greentechmedia.com. Retrieved 2019-11-29.
  5. "Hydrogen isn't the fuel of the future. It's already here". World Economic Forum. Retrieved 2019-11-29.
  6. Snyder, John (2019-09-05). "Hydrogen fuel cells gain momentum in maritime sector". Riviera Maritime Media.
  7. "Global Hydrogen Generation Market Size | Industry Report, 2027".
  8. UKCCC H2 2018, p. 20
  9. "Hydrogen could help decarbonise the global economy". Financial Times. Retrieved 2019-08-31.
  10. IEA H2 2019, p. 18
  11. National Hydrogen Association; United States Department of Energy. "The History of Hydrogen" (PDF). hydrogenassociation.org. National Hydrogen Association. p. 1. Archived from the original (PDF) on 14 July 2010. Retrieved 17 December 2010.
  12. Daedalus or Science and the Future, A paper read to the Heretics, Cambridge, on February 4th, 1923 Transcript 1993
  13. Jones, Lawrence W (13 March 1970). Toward a liquid hydrogen fuel economy. University of Michigan Environmental Action for Survival Teach In. Ann Arbor, Michigan: University of Michigan. hdl:2027.42/5800.
  14. Bakker, Sjoerd (2010). "The car industry and the blow-out of the hydrogen hype" (PDF). Energy Policy. 38 (11): 6540–6544. doi:10.1016/j.enpol.2010.07.019.
  15. Harrison, James. "Reactions: Hydrogen hype". Chemical Engineer. 58: 774–775.
  16. Rizzi, Francesco Annunziata, Eleonora Liberati, Guglielmo Frey, Marco (2014). "Technological trajectories in the automotive industry: are hydrogen technologies still a possibility?". Journal of Cleaner Production. 66: 328–336. doi:10.1016/j.jclepro.2013.11.069.CS1 maint: multiple names: authors list (link)
  17. Murai, Shusuke (2018-03-05). "Japan's top auto and energy firms tie up to promote development of hydrogen stations". The Japan Times Online. Japan Times. Retrieved 16 April 2018.
  18. Mishra, Ankit (2018-03-29). "Prospects of fuel-cell electric vehicles boosted with Chinese backing". Energy Post. Retrieved 16 April 2018.
  19. IEA H2 2019, p. 17
  20. IEA H2 2019, p. 14
  21. Crabtree, George W.; Dresselhaus, Mildred S.; Buchanan, Michelle V. (2004). The Hydrogen Economy (PDF) (Technical report).
  22. Mealey, Rachel. ”Automotive hydrogen membranes-huge breakthrough for cars", ABC, August 8, 2018
  23. "Archived copy". Argonne National Laboratory. Archived from the original on 2007-09-22. Retrieved 2007-06-15.CS1 maint: archived copy as title (link)
  24. Argonne National Laboratory. "Configuration and Technology Implications of Potential Nuclear Hydrogen System Applications" (PDF). Archived from the original (PDF) on 5 August 2013. Retrieved 29 May 2013.
  25. "Vehicle Technologies Program: Fact #205: February 25, 2002 Hydrogen Cost and Worldwide Production". .eere.energy.gov. Retrieved 2009-09-19.
  26. "Bellona-HydrogenReport". Interstatetraveler.us. Retrieved 2010-07-05.
  27. "Definition of Green Hydrogen" (PDF). Clean Energy Partnership. Retrieved 2014-09-06.
  28. "Brown coal the hydrogen economy stepping stone | ECT". Retrieved 2019-06-03.
  29. Sampson2019-02-11T10:48:00+00:00, Joanna. "Blue hydrogen for a green future". gasworld. Retrieved 2019-06-03.
  30. "Actual Worldwide Hydrogen Production from …". Arno A Evers. December 2008. Archived from the original on 2015-02-02. Retrieved 2008-05-09.
  31. IEA H2 2019, p. 37
  32. "How Much Electricity/Water Is Needed to Produce 1 kg of H2 by Electrolysis?". Retrieved 17 June 2020.
  33. Badwal, SPS (2014). "Emerging electrochemical energy conversion and storage technologies". Frontiers in Chemistry. 2: 79. Bibcode:2014FrCh....2...79B. doi:10.3389/fchem.2014.00079. PMC 4174133. PMID 25309898.
  34. Werner Zittel; Reinhold Wurster (1996-07-08). "Chapter 3: Production of Hydrogen. Part 4: Production from electricity by means of electrolysis". HyWeb: Knowledge - Hydrogen in the Energy Sector. Ludwig-Bölkow-Systemtechnik GmbH.
  35. Bjørnar Kruse; Sondre Grinna; Cato Buch (2002-02-13). "Hydrogen—Status and Possibilities". The Bellona Foundation. Archived from the original (PDF) on 2011-07-02. Efficiency factors for PEM electrolysers up to 94% are predicted, but this is only theoretical at this time.
  36. "high-rate and high efficiency 3D water electrolysis". Grid-shift.com. Archived from the original on 2012-03-22. Retrieved 2011-12-13.
  37. "Wide Spread Adaption of Competitive Hydrogen Solution" (PDF). http://nelhydrogen.com. Nel ASA. Retrieved 22 April 2018. External link in |website= (help)
  38. Philibert, Cédric. "Commentary: Producing industrial hydrogen from renewable energy". https://www.iea.org. International Energy Agency. Retrieved 22 April 2018. External link in |website= (help)
  39. https://www.hfpeurope.org/infotools/energyinfos__e/hydrogen/main03.html%5B%5D
  40. "High hydrogen yield from a two-step process of dark-and photo-fermentation of sucrose". Cat.inist.fr. Retrieved 2010-07-05.
  41. "Hydrogen production from organic solid matter". Biohydrogen.nl. Retrieved 2010-07-05.
  42. Hemschemeier, A; Melis, A; Happe, T (2009). "Analytical approaches to photobiological hydrogen production in unicellular green algae". Photosyn. Res. 102 (2–3): 523–40. doi:10.1007/s11120-009-9415-5. PMC 2777220. PMID 19291418.
  43. "NanoLogix generates energy on-site with bioreactor-produced hydrogen". Solid State Technology. September 20, 2007. Archived from the original on 2018-05-15. Retrieved 14 May 2018.
  44. "Power from plants using microbial fuel cell" (in Dutch). Retrieved 2010-07-05.
  45. "2001-High pressure electrolysis - The key technology for efficient H.2" (PDF). Retrieved 2010-07-05.
  46. Carmo, M; Fritz D; Mergel J; Stolten D (2013). "A comprehensive review on PEM water electrolysis". Journal of Hydrogen Energy. 38 (12): 4901–4934. doi:10.1016/j.ijhydene.2013.01.151.
  47. "2003-PHOEBUS-Pag.9" (PDF). Archived from the original (PDF) on 2009-03-27. Retrieved 2010-07-05.
  48. Finland exporting TEN-T fuel stations
  49. "Steam heat: researchers gear up for full-scale hydrogen plant" (Press release). Science Daily. 2008-09-18. Retrieved 2008-09-19.
  50. "Nuclear Hydrogen R&D Plan" (PDF). U.S. Dept. of Energy. March 2004. Archived from the original (PDF) on 2008-05-18. Retrieved 2008-05-09.
  51. Valenti G, Boni A, Melchionna M, Cargnello M, Nasi L, Bertoni G, Gorte R, Marcaccio M, Rapino S, Bonchio M, Fornasiero P, Prato M, Paolucci F (2016). "Co-axial heterostructures integrating palladium/ titanium dioxide with carbon nanotubes for efficient electrocatalytic hydrogen evolution". Nature Communications. 7: 13549. Bibcode:2016NatCo...713549V. doi:10.1038/ncomms13549. PMC 5159813. PMID 27941752.
  52. William Ayers, US Patent 4,466,869 Photolytic Production of Hydrogen
  53. del Valle, F.; Álvarez Galván, M. Consuelo; Del Valle, F.; Villoria De La Mano, José A.; Fierro, José L. G.; et al. (Jun 2009). "Water Splitting on Semiconductor Catalysts under Visible-Light Irradiation". ChemSusChem. 2 (6): 471–485. doi:10.1002/cssc.200900018. PMID 19536754.
  54. del Valle, F.; Del Valle, F.; Villoria De La Mano, J.A.; Álvarez-Galván, M.C.; Fierro, J.L.G.; et al. (2009). Photocatalytic water splitting under visible Light: concept and materials requirements. Advances in Chemical Engineering. 36. pp. 111–143. doi:10.1016/S0065-2377(09)00404-9. ISBN 9780123747631.
  55. Nann, Thomas; Ibrahim, Saad K.; Woi, Pei-Meng; Xu, Shu; Ziegler, Jan; Pickett, Christopher J. (2010-02-22). "Water Splitting by Visible Light: A Nanophotocathode for Hydrogen Production". Angewandte Chemie International Edition. 49 (9): 1574–1577. doi:10.1002/anie.200906262. PMID 20140925. Retrieved 2011-12-13.
  56. Yamamura, Tetsushi (August 2, 2015). "Panasonic moves closer to home energy self-sufficiency with fuel cells". Asahi Shimbun. Archived from the original on August 7, 2015. Retrieved 2015-08-02.
  57. "DLR Portal - DLR scientists achieve solar hydrogen production in a 100-kilowatt pilot plant". Dlr.de. 2008-11-25. Retrieved 2009-09-19.
  58. "353 Thermochemical cycles" (PDF). Retrieved 2010-07-05.
  59. UNLV Thermochemical cycle automated scoring database (public)
  60. "Development of Solar-powered Thermochemical Production of Hydrogen from Water" (PDF). Retrieved 2010-07-05.
  61. http://www.nedstack.com/images/stories/news/documents/20120202_Press%20release%20Solvay%20PEM%20Power%20Plant%20start%20up.pdf Archived 2014-12-08 at the Wayback Machine Nedstack
  62. Zubrin, Robert (2007). Energy Victory. Amherst, New York: Prometheus Books. pp. 117–118. ISBN 978-1-59102-591-7. The situation is much worse than this, however, because before the hydrogen can be transported anywhere, it needs to be either compressed or liquefied. To liquefy it, it must be refrigerated down to a temperature of -253°C (20 degrees above absolute zero). At these temperatures, fundamental laws of thermodynamics make refrigerators extremely inefficient. As a result, about 40 percent of the energy in the hydrogen must be spent to liquefy it. This reduces the actual net energy content of our product fuel to 792 kcal. In addition, because it is a cryogenic liquid, still more energy could be expected to be lost as the hydrogen boils away as it is warmed by heat leaking in from the outside environment during transport and storage.
  63. Savvides, Nick (2017-01-11). "Japan plans to use imported liquefied hydrogen to fuel Tokyo 2020 Olympics". https://fairplay.ihs.com/. IHS Markit Maritime Portal. Retrieved 22 April 2018. External link in |website= (help)
  64. S.Sadaghiani, Mirhadi (2 March 2017). "Introducing and energy analysis of a novel cryogenic hydrogen liquefaction process configuration". International Journal of Hydrogen Energy. 42 (9).
  65. 1994 – ECN abstract. Hyweb.de. Retrieved on 2012-01-08.
  66. European Renewable Energy Network pp. 86, 188
  67. "Energy storage – the role of electricity" (PDF). https://ec.europa.eu/. European Commission. Retrieved 22 April 2018. External link in |website= (help)
  68. "Hyunder".
  69. Storing renewable energy: Is hydrogen a viable solution?
  70. "BRINGING NORTH SEA ENERGY ASHORE EFFICIENTLY" (PDF). worldenergy.org. World Energy Council Netherlands. Retrieved 22 April 2018.
  71. GERDES, JUSTIN (2018-04-10). "Enlisting Abandoned Oil and Gas Wells as 'Electron Reserves'". https://www.greentechmedia.com. Wood MacKenzie. Retrieved 22 April 2018. External link in |website= (help)
  72. Anscombe, Nadya (4 June 2012). "Energy storage: Could hydrogen be the answer?". Solar Novus Today. Retrieved 3 November 2012.
  73. Naturalhy Archived 2012-01-18 at the Wayback Machine
  74. Kijk magazine, 10, 2019
  75. 50% hydrogen for Europe. A manifesto by Frank Wouters and Ad van Wijk
  76. Idaho National Engineering Laboratory's recommendation for Gaseous Hydrogen: Stainless steel Archived 2012-09-16 at Archive-It Accessed 2010-10-13
  77. Stuart Island Energy Initiative Website Accessed 2010-10-13: Hydrogen has an active electron, and therefore behaves somewhat like a Halogen. The recommended pipe material is stainless steel.
  78. IEA H2 2019, p. 15
  79. "Japan's Hydrogen Strategy and Its Economic and Geopolitical Implications". Etudes de l'Ifri. Retrieved 9 February 2019.
  80. "South Korea's Hydrogen Economy Ambitions". The Diplomat. Retrieved 9 February 2019.
  81. "The world´s largest-class hydrogen production, Fukushima Hydrogen Energy Research Field (FH2R) now is completed at Namie town in Fukushima". Toshiba Energy Press Releases. Toshiba Energy Systems and Solutions Corporations. 7 March 2020. Retrieved 1 April 2020.
  82. Editor (2019-06-14). "Hydrogen could replace natural gas to heat homes and slash carbon emissions, new report claims | Envirotec". Retrieved 2019-09-25.CS1 maint: extra text: authors list (link)
  83. Murray, Jessica (2020-01-24). "Zero-carbon hydrogen injected into gas grid for first time in groundbreaking UK trial". The Guardian. ISSN 0261-3077. Retrieved 2020-01-24.
  84. frankwouters1 (2019-05-07). "A European Hydrogen Manifesto". Frank Wouters. Retrieved 2019-12-02.
  85. "idealhy.eu - Liquid Hydrogen Outline". idealhy.eu. Retrieved 2019-12-02.
  86. Electricity from wood through the combination of gasification and solid oxide fuel cells, Ph.D. Thesis by Florian Nagel, Swiss Federal Institute of Technology Zurich, 2008
  87. "Power-to-weight ratio". .eere.energy.gov. 2009-06-23. Archived from the original on 2010-06-09. Retrieved 2010-07-05.
  88. "EPA mileage estimates". Honda FCX Clarity - Vehicle Specifications. American Honda Motor Company. Retrieved 17 December 2010.
  89. "Fuel Cell Technologies Office; Accomplishments and Progress". US Department of Energy. Retrieved 16 April 2018.
  90. R&D of large stationary hydrogen/CNG/HCNG storage vessels
  91. Kreith, 2004
  92. Seba, Tony (23 October 2015). "Toyota vs Tesla - hydrogen fuel cell vehicles vs electric cars". EnergyPost.eu. Retrieved 3 December 2016.
  93. Bossel, Ulrich (2006). "Does a Hydrogen Economy Make Sense?". Proceedings of the IEEE. 94 (10): 1826–1837. doi:10.1109/JPROC.2006.883715. Mirror
  94. Ann Mari Svensson; Steffen Møller-Holst; Ronny Glöckner; Ola Maurstad (September 2006). "Well-to-wheel study of passenger vehicles in the Norwegian energy system". Energy. 32 (4): 437–45. doi:10.1016/j.energy.2006.07.029.
  95. Boyd, Robert S. (May 15, 2007). "Hydrogen cars may be a long time coming". McClatchy Newspapers. Archived from the original on May 1, 2009. Retrieved 2008-05-09.
  96. Squatriglia, Chuck (May 12, 2008). "Hydrogen Cars Won't Make a Difference for 40 Years". Wired. Retrieved 2008-05-13.
  97. National Academy of Engineering (2004). The Hydrogen Economy: Opportunities, Costs, Barriers, and R&D Needs. Washington, D.C.: The National Academies Press. doi:10.17226/10922. ISBN 978-0-309-53068-2. Retrieved 17 December 2010.
  98. "Ford Motor Company Business Plan", December 2, 2008
  99. Dennis, Lyle. "Nissan Swears Off Hydrogen and Will Only Build Electric Cars", All Cars Electric, February 26, 2009
  100. Letter of Understanding 2009
  101. "Hydrogen for transport", The Carbon Trust, 28 November 2014. Retrieved on 20 January 2015.
  102. BMW Group Clean Energy ZEV Symposium. September 2006, p. 12
  103. "This company may have solved one of the hardest problems in clean energy". Vox. 2018-02-16. Retrieved 9 February 2019.
  104. Utgikar, Vivek P; Thiesen, Todd (2005). "Safety of compressed hydrogen fuel tanks: Leakage from stationary vehicles". Technology in Society. 27 (3): 315–320. doi:10.1016/j.techsoc.2005.04.005.
  105. "Hydrogen Sensor: Fast, Sensitive, Reliable, and Inexpensive to Produce" (PDF). Argonne National Laboratory. September 2006. Retrieved 2008-05-09.
  106. "Canadian Hydrogen Safety Program testing H2/CNG". Hydrogenandfuelcellsafety.info. Archived from the original on 2011-07-21. Retrieved 2010-07-05.
  107. UKCCC H2 2018, p. 113
  108. "A wake-up call on green hydrogen: the amount of wind and solar needed is immense | Recharge". Recharge | Latest renewable energy news. Retrieved 2020-04-11.
  109. UKCCC H2 2018, p. 7
  110. UKCCC H2 2018, p. 124
  111. UKCCC H2 2018, p. 118
  112. "Experimental 'wind to hydrogen' system up and running". Physorg.com. January 8, 2007. Retrieved 2008-05-09.
  113. "Hydrogen Engine Center Receives Order for Hydrogen Power Generator 250kW Generator for Wind/Hydrogen Demonstration" (PDF). Hydrogen Engine Center, Inc. May 16, 2006. Archived from the original (PDF) on May 27, 2008. Retrieved 2008-05-09.
  114. "Stuart Island Energy Initiative". Retrieved 2008-05-09.
  115. Hydrogen transport & distribution
  116. "Hydrogen buses". Transport for London. Archived from the original on March 23, 2008. Retrieved 2008-05-09.
  117. "The Hydrogen Expedition" (PDF). January 2005. Archived from the original (PDF) on 2008-05-27. Retrieved 2008-05-09.
  118. "Perth Fuel Cell Bus Trial". Department for Planning and Infrastructure, Government of Western Australia. 13 April 2007. Archived from the original on 7 June 2008. Retrieved 2008-05-09.
  119. Hannesson, Hjálmar W. (2007-08-02). "Climate change as a global challenge". Iceland Ministry for Foreign Affairs. Retrieved 2008-05-09.
  120. Doyle, Alister (January 14, 2005). "Iceland's hydrogen buses zip toward oil-free economy". Reuters. Archived from the original on July 24, 2012. Retrieved 2008-05-09.
  121. "What is HyFLEET:CUTE?". Archived from the original on 2008-02-24. Retrieved 2008-05-09.
  122. Hydrogen vehicles and refueling infrastructure in India
  123. L. M. DAS, EXHAUST EMISSION CHARACTERIZATION OF HYDROGEN OPERATED ENGINE SYSTEM: NATURE OF POLLUTANTS AND THEIR CONTROL TECHNIQUES Int. J. Hydrogen Energy Vol. 16, No. 11, pp. 765-775, 1991
  124. MNRE: FAQ
  125. Overview of Indian Hydrogen Programme
  126. H2 stations worldwide
  127. India working on more H2 stations
  128. Shell plans to open 1200 fuel stations in India, some of which may include H2 refilling
  129. Hydrogen Vehicles and Refueling Infrastructure in India
  130. "Independent Mid-Term Review of the UNIDO Project: Establishment and operation of the International Centre for Hydrogen Energy Technologies (ICHET), TF/INT/03/002" (PDF). UNIDO. 31 August 2009. Archived from the original (PDF) on 1 June 2010. Retrieved 20 July 2010.
  131. Worldwide NGV statistics
  132. Fuel Cell micro CHP
  133. Fuel cell micro Cogeneration
  134. Agosta, Vito (July 10, 2003). "The Ammonia Economy". Archived from the original on May 13, 2008. Retrieved 2008-05-09.
  135. "Renewable Energy". Iowa Energy Center. Archived from the original on 2008-05-13. Retrieved 2008-05-09.
  136. UKCCC H2 2018, p. 36: "Near-term pursuit of hybrid heat pumps would not necessarily lead to a long-term solution of hybrid heat pumps with hydrogen boilers."
  137. UKCCC H2 2018, p. 79: The potential for bio-gasification with CCS to be deployed at scale is limited by the amount of sustainable bioenergy available. .... "
  138. UKCCC H2 2018, p. 33: production of biofuels, even with CCS, is only one of the best uses of the finite sustainable bio-resource if the fossil fuels it displaces cannot otherwise feasibly be displaced (e.g. use of biomass to produce aviation biofuels with CCS)."

Sources

This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.