Sei sulla pagina 1di 9

Hydrogen Production, Storage & Safety

(Edited from http://www.eere.energy.gov/hydrogenandfuelcells/production/)

Hydrogen can be produced using diverse, domestic resources including fossil fuels, such as natural gas and coal
(with carbon sequestration); nuclear; and biomass and other renewable energy technologies, such as wind, solar,
geothermal, and hydro-electric power. The overall challenge to hydrogen production is cost reduction. For
transportation, a key driver for energy independence and therefore the hydrogen economy, hydrogen must be
cost-competitive with conventional fuels and technologies on a per-mile basis in order to succeed in the
commercial marketplace.

1. Basics
Hydrogen, chemical symbol H, is the simplest element on earth. An atom of hydrogen has only one proton and
one electron. Hydrogen gas is a diatomic molecule - each molecule has two atoms of hydrogen. Although
abundant on earth as an element, hydrogen combines readily with other elements and is almost always found as
part of some other substance, such as water, hydrocarbons, or alcohols. It is also found in biomass, which
includes all plants and animals.

• Hydrogen is an energy carrier, not an energy source. Hydrogen can store and deliver usable energy, but it
doesn't typically exist by itself in nature; it must be produced from compounds that contain it.
• Hydrogen can be produced using diverse, domestic resources including nuclear; natural gas and coal; and
biomass and other renewables including solar, wind, hydroelectric, or geothermal energy. This diversity of
domestic energy sources makes hydrogen a promising energy carrier and important to energy security. It is
expected and desirable for hydrogen to be produced using a variety of resources and process technologies
(or pathways).
• It is necessary to focus on hydrogen production technologies that result in near-zero net greenhouse gas
emissions and use renewable energy sources, nuclear energy, and coal (when combined with carbon
sequestration). To ensure sufficient clean energy for our overall energy needs, energy efficiency is also
important.
• Hydrogen can be produced via various process technologies, including thermal (natural gas reforming,
renewable liquid and bio-oil processing, and biomass and coal gasification), electrolytic (water splitting
using a variety of energy resources), and photolytic (splitting water using sunlight via biological and
electrochemical materials).

2. Hydrogen Production Technology


The development of clean, sustainable, and cost-competitive hydrogen production processes are key to a viable
future hydrogen economy. Hydrogen production technologies fall into three general categories:
1. Thermal Processes
2. Electrolytic Processes
3. Photolytic Processes

2.1 Thermal Processes


Some thermal processes use the energy in various resources, such as natural gas, coal, or biomass, to release
hydrogen, which is part of their molecular structure. In other processes, heat, in combination with closed
chemical cycles, produces hydrogen from feedstocks such as water - these are known as "thermochemical"
processes.
1. Reforming of Natural Gas
2. Gasification of Coal
3. Gasification of Biomass
4. Reforming of Renewable Liquid Fuels
5. High-temperature Water Splitting

MEE507 1
2.1.1 Natural Gas Reforming
Natural gas reforming is an important pathway for near-term hydrogen production. Today most hydrogen is
produced from fossil materials, such as from natural gas at oil refinery. Natural gas contains methane (CH4) that
can be used to produce hydrogen via thermal processes, such as steam methane reformation and partial
oxidation.
2.1.1.1 Steam Methane Reforming
About 95% of the hydrogen produced today in the US is made via steam methane reforming, a process in which
high-temperature steam (700 - 1000°C) is used to produce hydrogen from natural gas. In steam methane
reforming, methane reacts with steam under 3-25 bar pressure in the presence of a catalyst to produce hydrogen,
carbon monoxide, and a relatively small amount of carbon dioxide. Steam reforming is endothermic.
Subsequently, in what is called the water-gas shift reaction, the carbon monoxide and steam are reacted using a
catalyst to produce carbon dioxide and more hydrogen. In a final process step called pressure-swing adsorption,
carbon dioxide and other impurities are removed from the gas stream, leaving essentially pure hydrogen. Steam
reforming can also be used to produce hydrogen from other fuels, such as ethanol, propane, or even gasoline.
Steam Reforming Reactions
Methane: CH4 + H2O (+heat) → CO + 3H2
Propane: C3H8 + 3H2O (+heat) → 3CO + 7H2
Ethanol: C2H5OH + H2O (+heat) → 2CO + 4H2
Water-Gas Shift Reaction
CO + H2O → CO2 + H2 (+small amount of heat)

2.1.1.2 Partial Oxidation


In partial oxidation, the methane and other hydrocarbons in natural gas are reacted with a limited amount of
oxygen (typically, from air) that is not enough to completely oxidize the hydrocarbons to carbon dioxide and
water. With less than the stoichiometric amount of oxygen available for the reaction, the reaction products
contain primarily hydrogen and carbon monoxide (and nitrogen, if the reaction is carried out with air rather than
pure oxygen), and a relatively small amount of carbon dioxide and other compounds. Subsequently, in a water-
gas shift reaction, the carbon monoxide reacts with water to form carbon dioxide and more hydrogen.
Partial Oxidation Reactions
Methane: CH4 + ½O2 → CO + 2H2 (+heat)
Propane: C3H8 + 1½O2 → 3CO + 4H2 (+heat)
Ethanol: C2H5OH + ½O2 → 2CO + 3H2 (+heat)
Water-Gas Shift Reaction
CO + H2O → CO2 + H2 (+small amount of heat)
Producing hydrogen from natural gas does result in some greenhouse gas emissions. When compared to internal
combustion engine vehicles using gasoline, however, fuel cell vehicles using hydrogen produced from natural
gas reduce greenhouse gas emissions by 60%.

2.1.2 Coal Gasification


Coal is converted into a gaseous mixture of hydrogen, carbon monoxide, carbon dioxide, and other compounds
by applying heat under pressure in the presence of steam and a controlled amount of oxygen in a gasifier. The
coal is chemically broken apart by the gasifier's heat, steam, and oxygen, setting into motion chemical reactions
that produce a synthesis gas, or syngas - a mixture of primarily hydrogen, carbon monoxide, and carbon dioxide.
The carbon monoxide is reacted (in a separate unit) with water to form carbon dioxide and more hydrogen.
Adsorbers or special membranes can separate the hydrogen from this gas stream.
Chemically, coal is a complex and highly variable substance. The carbon and hydrogen in coal may be
represented in approximate manner as 0.8 atoms of hydrogen per atom of carbon in bituminous coal. Its
gasification reaction may be represented by the (unbalanced) reaction equation:
CH0.8 + O2 + H2O → CO + CO2 + H2 + other species
An advantage of this technology is that carbon dioxide can be separated more easily from the syngas and
captured, instead of being released into the atmosphere. If carbon dioxide can be successfully sequestered,
hydrogen can be produced from coal gasification with near-zero greenhouse gas emissions. Coal gasification

MEE507 2
can also be used to produce electricity by routing the syngas to a turbine to generate electricity. Coal gasification
technology could be used to generate both electricity and hydrogen in one integrated plant operation. Coal
gasification technology is most appropriate for large-scale, centralized hydrogen production.

2.1.3 Biomass Gasification


Biomass is converted into a gaseous mixture of hydrogen, carbon monoxide, carbon dioxide, and other
compounds by applying heat under pressure in the presence of steam and a controlled amount of oxygen in a
gasifier. The biomass is chemically broken apart by the gasifier's heat, steam, and oxygen, setting into motion
chemical reactions that produce a synthesis gas, or "syngas"—a mixture of primarily hydrogen, carbon
monoxide, and carbon dioxide. The carbon monoxide is then reacted with water to form carbon dioxide and
more hydrogen (water-gas shift reaction). Adsorbers or special membranes can separate the hydrogen from this
gas stream.
Simplified Reaction
C6H12O6 + O2 + H2O → CO + CO2 + H2 + other species
Water-Gas Shift Reaction
CO + H2O → CO2 + H2 (+small amount of heat)
Pyrolysis is the gasification of biomass in the absence of oxygen. In general, biomass does not gasify as easily
as coal, and it produces other hydrocarbon compounds in the gas mixture exiting the gasifier; this is especially
true when no oxygen is used. As a result, typically an extra step must be taken to reform these hydrocarbons
with a catalyst to yield a clean syngas mixture of hydrogen, carbon monoxide, and carbon dioxide. Then, just as
in the gasification process for hydrogen production, a shift reaction step (with steam) converts the carbon
monoxide to carbon dioxide. The hydrogen produced is then separated and purified.
Biomass gasification technology is most appropriate for large-scale, centralized hydrogen production, due to the
nature of handling large amounts of biomass and the required economy of scale for this type of process.
Biomass is an abundant natural and renewable resource. Growing biomass removes carbon dioxide from the
atmosphere. The biomass resources used in biomass gasification consume carbon dioxide in the atmosphere as
part of their natural growth process, which means that biomass gasification results in a near-zero net release of
greenhouse gases.

2.1.4 Renewable Liquid Fuels Reforming


Reforming renewable liquids like ethanol, bio-oils, or other liquid fuels to hydrogen is very similar to reforming
natural gas. The liquid fuel is reacted with steam at high temperatures in the presence of a catalyst to produce a
reformate gas composed mostly of hydrogen and carbon monoxide. Additional hydrogen and carbon dioxide are
produced by reacting the carbon monoxide (created in the first step) with high temperature steam in the "water-
gas shift reaction". Finally, the hydrogen is separated out and purified.
Steam Reforming Reaction (Ethanol)
C2H5OH + H2O (+heat) → 2CO + 4H2
Water-Gas Shift Reaction
CO + H2O → CO2 + H2 (+small amount of heat)

2.1.5 High-Temperature Water Splitting


High-temperature water splitting is a long-term technology in the early stages of development. High-
temperature heat (500 - 2000°C) drives a series of chemical reactions that produce hydrogen. Chemicals used in
the process are reused within each cycle, creating a closed loop that consumes only water and produces
hydrogen and oxygen. The high-temperature heat needed can be supplied by next-generation nuclear reactors
under development (up to about 1000°C) or by using sunlight with solar concentrators (up to about 2000°C).
High temperature water splitting is most suitable for large scale, centralized production of hydrogen, although
semi-central production from solar driven cycles might be possible. Solar- and nuclear-driven high-temperature
thermochemical water splitting cycles produce hydrogen with near-zero greenhouse gas emissions, using water and
either sunlight or nuclear energy.
A solar concentrator uses mirrors and a reflective or refractive lens to capture and focus sunlight to produce
temperatures up to 2,000°C. This high temperature heat can be used to drive chemical reactions that produce
hydrogen.

MEE507 3
Example: zinc/zinc oxide chemical cycle
Zinc oxide powder is passed through a reactor heated by a solar concentrator operating at about 1,900°C. At this
temperature, the zinc oxide dissociates to zinc and oxygen gases. The zinc is cooled, separated, and reacted with
water to form hydrogen gas and solid zinc oxide. The net result is hydrogen and oxygen, produced from water.
The hydrogen can be separated and purified. The zinc oxide can be recycled and reused to create more hydrogen
through this process.
2ZnO + heat → 2Zn + O2
2Zn + 2H2O → 2ZnO + 2H2

2.2 Electrolytic Processes


Electrolytic processes use electricity to split water into hydrogen and oxygen, a process that takes place in an
electrolyzer. Hydrogen produced via electrolysis can result in zero greenhouse gas emissions, depending on the source
of the electricity used. The two electrolysis pathways of greatest interest for wide-scale hydrogen production, which
result in near-zero greenhouse gas emissions, are electrolysis using renewable sources of electricity and nuclear high-
temperature electrolysis. Like fuel cells, electrolyzers consist of an anode and a cathode separated by an
electrolyte. Different electrolyzers function in slightly different ways.
2.2.1 PEM Electrolyzer
In a polymer electrolyte membrane (PEM) electrolyzer, the electrolyte is a solid specialty plastic material. Water
reacts at the anode to form oxygen and positively charged hydrogen ions (protons). The electrons flow through
an external circuit and the hydrogen ions selectively move across the PEM to the cathode. At the cathode,
hydrogen ions combine with electrons from the external circuit to form hydrogen gas.
Anode Reaction: 2H2O → O2 + 4H+ + 4e-
Cathode Reaction: 4H+ + 4e- → 2H2
2.2.2 Alkaline Electrolyzers
Alkaline electrolyzers are similar to PEM electrolyzers but use an alkaline solution (of sodium or potassium
hydroxide) that acts as the electrolyte. These electrolyzers have been commercially available for many years.
2.2.3 Solid Oxide Electrolyzers
Solid oxide electrolyzers, which use a solid ceramic material as the electrolyte that selectively transmits
negatively charged oxygen ions at elevated temperatures, generate hydrogen in a slightly different way. Water at
the cathode combines with electrons from the external circuit to form hydrogen gas and negatively charged
oxygen ions. The oxygen ions pass through the membrane and react at the anode to form oxygen gas and give
up the electrons to the external circuit. Solid oxide electrolyzers must operate at temperatures high enough for
the solid oxide membranes to function properly (about 500 - 800°C; compared to PEM electrolyzers, which
operate at 80 - 100°C, and alkaline electrolyzers, which operate at 100-150°C). The solid oxide electrolyzers can
effectively use heat available at these elevated temperatures from various sources, including renewable energy,
to decrease the amount of electrical energy needed to produce hydrogen from water.
Hydrogen production via electrolysis may offer opportunities for synergy with variable power generation, which
is characteristic of some renewable energy technologies. Hydrogen fuel and electric power generation could be
integrated at a wind farm, allowing flexibility to shift production to best match resource availability with system
operational needs and market factors.

2.3 Photolytic Processes


Photolytic processes use light energy to split water into hydrogen and oxygen.
Currently in the very early stages of research, these processes offer long-term
potential for sustainable hydrogen production with low environmental impact. The
two technological pathways considered are:
1. Photobiological Water Splitting
2. Photoelectrochemical Water Splitting

2.3.1 Photobiological Water Splitting


In this process, hydrogen is produced from water using sunlight and specialized
microorganisms, such as green algae and cyanobacteria. Just as plants produce
oxygen during photosynthesis, these microorganisms consume water and produce

MEE507 4
hydrogen as a byproduct of their natural metabolic processes. Photobiological water splitting is a long-term
technology. Currently, the microbes split water much too slowly to be used for efficient, commercial hydrogen
production. A set of bio-reactors use can use light (sunlight or artificial light) and the natural activities of
enzymes in green algae to produce hydrogen from water as seen in the Figure.

2.3.2 Photoelectrochemical Water Splitting


In this process, hydrogen is produced from water using sunlight and
specialized semiconductors called photoelectrochemical materials. In the
photoelectrochemical system, the semiconductor uses light energy to directly
dissociate water molecules into hydrogen and oxygen. Different
semiconductor materials work at particular wavelengths of light and energies.
Research focuses on finding semiconductors with the correct energies to split
water that are also stable when in contact with water. Photobiological water
splitting is in the very early stages of research. In the Figure, light shining on a
photoelectrochemical cell immersed in water produces bubbles of hydrogen
and oxygen.

3. Hydrogen Storage
Developing safe and reliable hydrogen storage technologies that meet performance and cost requirements is
critical to achieving a future hydrogen economy. Hydrogen storage will be needed for both vehicular
applications and off-board uses such as for stationary power generation and for hydrogen delivery and refueling
infrastructure.

3.1 Basics
Hydrogen can be physically stored as either a gas or a liquid. Storage as a gas typically requires high-pressure
tanks (5000-10,000 psi). Storage of hydrogen as a liquid requires cryogenic temperatures, since the boiling point
of hydrogen at one atmosphere pressure is -252.8ºC. Hydrogen can also be stored on the surfaces of solids (by
adsorption) or within solids (by absorption). In adsorption, hydrogen is attached to the surface of a material
(Figure) either as hydrogen molecules or as hydrogen atoms. In absorption, hydrogen is dissociated into H-
atoms and then the hydrogen atoms are incorporated into the solid lattice framework (Figure).

Hydrogen storage by absorption Hydrogen storage by adsorption

Hydrogen storage in solids may make it possible to store larger quantities of hydrogen in smaller volumes at low
pressure and at temperatures close to room temperature. It is also possible to achieve volumetric storage
densities greater than liquid hydrogen because the hydrogen molecule is dissociated into atomic hydrogen
within the metal hydride lattice structure. Hydrogen can be stored through the reaction of hydrogen-containing
materials with water or other compounds such as alcohols. In this case, the hydrogen is effectively stored in both
the material and in the water. The term chemical hydrogen storage or chemical hydride is used to describe this
form of hydrogen storage. It is also possible to store hydrogen in the chemical structures of liquids and solids.

3.2 Hydrogen Storage Technologies


Current on-board hydrogen storage approaches involve compressed hydrogen gas tanks, liquid hydrogen tanks,
metal hydrides and chemical hydrogen storage. Storage as a gas or liquid or storage in metal hydrides or high
surface area sorbents constitute reversible on-board hydrogen storage systems, since hydrogen regeneration or
refill can take place on-board the vehicle. For chemical hydrogen storage approaches such as a chemical

MEE507 5
reaction on board the vehicle to produce hydrogen, hydrogen regeneration is not possible on-board the vehicle
and thus these spent materials must be removed from the vehicle and regenerated off board.

3.2.1 Gaseous and Liquid Hydrogen Storage


Today's state-of-the-art for hydrogen storage includes 5000- and 10,000-psi compressed gas tanks and cryogenic
liquid hydrogen tanks for on-board hydrogen storage.

3.2.1.1 Compressed Hydrogen Gas Tanks


The energy density of gaseous hydrogen can be improved by storing hydrogen at higher pressures. This requires
material and design improvements in order to ensure tank integrity. Advances in compression technologies are
also required to improve efficiencies and reduce the cost of producing high-pressure hydrogen.
Carbon fiber-reinforced 5000-psi and 10,000-psi compressed hydrogen gas tanks are under development. Such
tanks are already in use in prototype hydrogen-powered vehicles. The inner liner of the tank is a high molecular
weight polymer that serves as a hydrogen gas permeation barrier. A carbon fiber-epoxy resin composite shell is
placed over the liner and constitutes the gas pressure load-bearing component of the tank. Finally, an outer shell
is placed on the tank for impact and damage resistance. The pressure regulator for the 10,000-psi tank is located
in the interior of the tank. There is also an in-tank gas temperature sensor to monitor the tank temperature during
the gas filling process when heating of the tank occurs.
One approach being pursued is to increase the gravimetric and volumetric storage capacities of compressed gas
tanks from their current levels. The first approach involves cryo-compressed tanks. This is based on the fact
that, at fixed pressure and volume, gas tank volumetric capacity increases as the tank temperature decreases.
Thus, by cooling a tank from room temperature to liquid nitrogen temperature (77 K), its volumetric capacity
will increase by a factor of four, although system volumetric capacity will be less than this due to the increased
volume required for the cooling system.

3.2.1.2 Liquid Hydrogen Tanks


The energy density of hydrogen can be improved by storing
hydrogen in a liquid state. However, the issues with LH2
tanks are hydrogen boil-off, the energy required for
hydrogen liquefaction, volume, weight, and tank cost. The
energy requirement for hydrogen liquefaction is high;
typically 30% of the heating value of hydrogen is required
for liquefaction. Hydrogen boil-off must be minimized or
eliminated for cost, efficiency and vehicle range
considerations, as well as for safety considerations when
vehicles are parked in confined spaces. Insulation is required
for LH2 tanks and this reduces system gravimetric and
volumetric capacity. Liquid hydrogen (LH2) tanks can store more hydrogen in a given volume than compressed
gas tanks. The volumetric capacity of liquid hydrogen is 0.070 kg/L, compared to 0.030 kg/L for 10,000 psi gas
tanks.

3.2.2 Materials-based Hydrogen Storage


There are presently three generic mechanisms known for storing hydrogen in materials: absorption, adsorption,
and chemical reaction.
Absorption: In absorptive hydrogen storage, hydrogen is absorbed directly into the bulk of the material. In
simple crystalline metal hydrides, this absorption occurs by the incorporation of atomic hydrogen into interstitial
sites in the crystallographic lattice structure.
Adsorption: Adsorption may be subdivided into physisorption and chemisorption, based on the energetics of
the adsorption mechanism. Physisorbed hydrogen is more weakly energetically bound to the material than is
chemisorbed hydrogen. Sorptive processes typically require highly porous materials to maximize the surface
area available for hydrogen sorption to occur, and to allow for easy uptake and release of hydrogen from the
material.
Chemical reaction: The chemical reaction route for hydrogen storage involves displacive chemical reactions
for both hydrogen generation and hydrogen storage. For reactions that may be reversible on-board a vehicle,
hydrogen generation and hydrogen storage take place by a simple reversal of the chemical reaction as a result of

MEE507 6
modest changes in the temperature and pressure. Sodium alanate-based complex metal hydrides are an example.
In many cases, the hydrogen generation reaction is not reversible under modest temperature/pressure changes.
Therefore, although hydrogen can be generated on-board the vehicle, getting hydrogen back into the starting
material must be done off-board. Sodium borohydride is an example.
Typical storage materials are:
Metal hydrides: reversible solid-state materials that can be regenerated on-board
Chemical hydrides: hydrogen is released via chemical reaction (usually with water); the “spent fuel”
or byproduct is regenerated off-board
Carbon-based materials: reversible solid-state materials that can be regenerated on-board

3.2.2.1 Metal Hydrides


Metal hydrides have the potential for reversible on-board hydrogen storage and release at low temperatures and
pressures. The optimum operating P-T window for
PEM fuel cell vehicular applications is in the range
of 1-10 atm and 25-120°C. A simple metal hydride
such as LaNi5H6, that incorporates hydrogen into
its crystal structure, can function in this range, but
its gravimetric capacity is too low (~1.3 wt.%) and
its cost is too high for vehicular applications.
Complex metal hydrides such as alanate (AlH4)
materials have the potential for higher gravimetric
hydrogen capacities in the operational window
than simple metal hydrides. Alanates can store and
release hydrogen reversibly when catalyzed with
titanium dopants, according to the following 2-step
displacive reaction for sodium alanate:
NaAlH4 = 1/3 Na3AlH6 +2/3Al+H2
Na3AlH6 = 3 NaH + Al + 3/2H2
The amount of hydrogen that a material can release, rather than only the amount the material can hold, is the key
parameter used to determine system (net) gravimetric and volumetric capacities. Issues with complex metal
hydrides include low hydrogen capacity, slow uptake and release kinetics, and cost.

3.2.2.2 Chemical Hydrogen Storage


The term chemical hydrogen storage is used to describe storage technologies in which hydrogen is generated
through a chemical reaction. Common reactions involve chemical hydrides with water or alcohols. Typically,
these reactions are not easily reversible on-board a vehicle. Hence, the 'spent fuel' and/or byproducts must be
removed from the vehicle and regenerated off-board.

Hydrolysis Reactions
Hydrolysis reactions involve the oxidation reaction of chemical hydrides with water to produce hydrogen. The
reaction of sodium borohydride has been the most studied to date. This reaction is as follows.
In the first embodiment, a slurry of an inert stabilizing liquid protects the hydride from contact with moisture
and makes the hydride pumpable. At the point of use, the slurry is mixed with water and the consequent reaction
produces high purity hydrogen.
NaBH4 + 2H2O = NaBO2 + 4H2
The reaction can be controlled in an aqueous medium via pH and the use of a catalyst. While the material
hydrogen capacity can be high and the hydrogen release kinetics fast, the borohydride regeneration reaction
must take place off-board.
Another hydrolysis reaction that is presently being investigated is the reaction of MgH2 with water to form
Mg(OH)2 and H2. In this case, particles of MgH2 are contained in a non-aqueous slurry to inhibit premature
water reactions when hydrogen generation is not required. Material-based capacities for the MgH2 slurry
reaction with water can be as high as 11 wt.%. However, similar to the sodium borohydride approach, water
must also be carried on-board the vehicle in addition to the slurry and the Mg(OH)2 must be regenerated off
board.

MEE507 7
Hydrogenation/Dehydrogenation Reactions
Hydrogenation and dehydrogenation reactions have been studied for many years as a means of hydrogen
storage. For example, the decalin-to-naphthalene reaction can release 7.3 wt.% hydrogen at 210ºC via the
reaction:
C10H18 = C10H8 + 5H2
A platinum-based or noble metal supported catalyst is required to enhance the kinetics of hydrogen evolution.

4. Hydrogen Safety
Hydrogen, in vast quantities, has been used safely in chemical and metallurgical applications, the food industry,
and the space program for many years. As hydrogen and fuel cells begin to play a greater role in meeting the
energy needs of our nation and the world, minimizing the safety hazards related to the use of hydrogen as a fuel
is essential.
Hydrogen has a long history of safe use in the chemical and aerospace industries. An understanding of hydrogen
properties, proper safety precautions and engineering controls, and established rules, regulations and standards
are the keys to this successful track record. As the use of hydrogen and fuel cell systems expand, codes and
standards will be needed to provide the information to safely build, maintain, and operate hydrogen and fuel cell
systems and facilities, to ensure uniformity of safety requirements, and to assure local code officials and safety
inspectors that sufficient safety standards have been met.
With proper handling and controls, hydrogen can be as safe as, or safer than, other fuels we use today. Safety
considerations associated with handling hydrogen include fire, explosion, and asphyxiation. Below is a chart
that shows how hydrogen stacks up against some common fuels.
Properties of Hydrogen, Natural Gas, Gasoline, and Propane
Hydrogen Natural Gas Gasoline Propane
(gas) (gas) (liquid) (liquid)
Lower heating value (BTU/lb) 51,532 21,300 18,000 - 19,000 19,800
Density at standard conditions 0.0007a 0.005a 6.0-6.5a 4.22
(pounds per gallon)
Autoignition temperature in air 1,050 - 1,080 1,004 495 850 - 950
(°F)
Volume concentrations for 4.1 - 74 5.3 - 15 1.4 - 7.6 2.2 - 9.5
flammability in air (%)
Diffusion coefficient in air 0.0946b 0.0248b 0.008b 0.017c
(inches squared/second)
Toxicity to humans Non-toxic, simple Non-toxic, simple Poisonous, Non-toxic,
asphyxiant asphyxiant irritant to lungs, simple
stomach and skin asphyxiant

4.1 Safe operating practices


Safe operating practices are established to minimize the known hazards associated with handling hydrogen -
fire, explosion, and asphyxiation. Some hazards can be mitigated by hydrogen's unique properties. For example,
hydrogen's high dispersion coefficient allows it to dissipate rapidly and makes it virtually impossible for
hydrogen to explode in an open area. Other hazards are minimized through operator training and proper system
design. Some representative examples include:

• Purging hydrogen systems with an inert gas such as nitrogen is required to avoid the formation of
flammable hydrogen/oxygen mixtures.
• Adequate ventilation can minimize eliminate the potential hazard of asphyxiation and the formation of
combustible hydrogen/oxygen mixtures.
• Because hydrogen burns with an almost invisible blue flame, special flame detectors are required.

MEE507 8
4.2 Facts of Hydrogen Safety
• As the lightest and smallest element in the universe, confining hydrogen is very difficult. Hydrogen is much
lighter than air and rises at a speed of almost 20 meters per second — two times faster than helium and six
times faster than natural gas — which means that when released, it rises and disperses quickly.
• Combustion cannot occur in a tank or any contained location that contains only hydrogen. An oxidizer, such
as oxygen, must be present.
• Hydrogen is odorless, colorless, and tasteless and therefore undetectable by human senses. For these and
other reasons, industry designs systems with ventilation and leak detection. Natural gas is also odorless,
colorless, and tasteless, but industry adds a sulfur-containing odorant so people can detect it. These odorants
are not used with hydrogen, however, because there is no known odorant light enough to “travel with”
hydrogen, and at the same dispersion rate. Current odorants also contaminate fuel cells, a popular hydrogen
application.
• Hydrogen burns very quickly. Under optimal combustion conditions, the energy required to initiate
hydrogen combustion is significantly lower than that required for other common fuels, such as natural gas
or gasoline. At low concentrations of hydrogen fuel in air, the energy required to initiate combustion is
similar to that of other fuels.
• Hydrogen flames have low radiant heat. A hydrogen fire has significantly less radiant heat when compared
to a hydrocarbon fire. Since low levels of heat are emitted near a hydrogen flame (the flame itself is just as
hot), the risk of secondary fires is lower.
• With the exception of oxygen, any gas can cause asphyxiation in high enough concentrations. In most
scenarios, however, because hydrogen rises and disperses so rapidly, it is unlikely to be confined where
asphyxiation might otherwise occur.
• Hydrogen is non-toxic and non-poisonous. It will not contaminate groundwater (it’s a gas under normal
atmospheric conditions), and a release of hydrogen is not known to contribute to atmospheric pollution or
water pollution.

MEE507 9

Potrebbero piacerti anche