Sunteți pe pagina 1din 34

Tyndall˚Centre

for Climate Change Research

Hydrogen Energy Technology

Geoff Dutton

April 2002

Tyndall Centre for Climate Change Research Working Paper 17


Hydrogen Energy Technology

Dr Geoff Dutton
Energy Research Unit (ERU)
CLRC Rutherford Appleton Laboratory
Chilton, Didcot, Oxon. OX11 0QX
Tel: 01235 445823 Fax: 01235 446863
Email: a.g.dutton@rl.ac.uk

Tyndall Working Paper No. 17


April 2002

Page 1
Contents

Introduction
Hydrogen
Hydrogen production
Steam methane reforming (SMR)
Partial oxidation
Integrated gasification combined cycle (IGCC)
Pyrolysis
Water electrolysis
Reversible fuel cell
Hydrogen bromide electrolysis
Hydrogen production during manufacture of chlorine
Photoelectrolysis
Biological hydrogen production
Hydrogen storage and distribution
Compressed gas
Liquefaction
Solid state hydrogen storage
Metal hydride storage systems
Hydride hydrolysis
Glass microspheres
Relative merits of hydrogen storage systems and comparison of costs
Hydrogen distribution and transport
Hydrogen end use systems
Hydrogen-fuelled internal combustion engines
Hydrogen-fuelled turbines
Fuel cells
Hydrogen systems
Overall economics of hydrogen systems
Conclusions
Appendix 1: Energy units
Appendix 2: Hydrogen characteristics and safety
Appendix 3: Basic format for technology summaries
References

Page 2
Introduction
The term hydrogen economy was first used during the energy crises of the 1970s to describe a national (or
international) energy infrastructure based on hydrogen produced from non-fossil primary energy sources.
Within this concept, hydrogen is regarded as a suitable storage and transmission vector for energy from
renewable or nuclear power systems, allowing the generator or utility increased flexibility in responding to
fluctuations in wind or solar input or consumer demand, on a short term (minutes/hours) or seasonal basis.
Hydrogen can be stored and transported in pressure vessels or transmitted by pipeline to the point of end-
use. It is a versatile fuel, which can easily be substituted for traditional fuels, whether for stationary or
transport applications, resulting in improved efficiency and negligible pollution at the point of use.

Considering the current need to develop responses to human-induced climate change, a fully developed
hydrogen economy has the potential to drastically reduce emissions of carbon dioxide within the electrical
power, transport, and low grade heat supply sectors. However, the energy path from solar, wind, and other
renewable generators, through hydrogen production via electrolysis and widespread storage and
distribution, to end-use in cars, aeroplanes, and domestic and business premises is complex and currently
very expensive. Intermediate paths, employing hydrogen derived from fossil fuel sources, are already used
to produce merchant hydrogen and are likely to be more economic (subject to fluctuations in fossil fuel
price) in the short to medium term. Most, but not all, of these fossil fuel based paths will require potentially
expensive carbon dioxide sequestration, if they are to contribute to the reduction of greenhouse gases.

This report on hydrogen energy conversion technologies aims to identify the current state of the art in terms
of typical plant sizes, readiness for large scale application, estimated capital and running costs, and the need
and potential for significant innovation against time scales of 10, 20, and 50 years.

The report has been prepared as the first of a series of three documents being prepared with funding from
the Tyndall Centre for Climate Change, as Phase I of the project: The Hydrogen Energy Economy: its long
term role in greenhouse gas reduction. The other two reports are concerned with lessons to be learned for
hydrogen energy development from previous transitions in large technical systems and energy futures
scenarios, and the potential use of hydrogen within transport systems. It has been prepared by consulting
many sources, including presentations made by leading industrial and academic experts to the UK
Hydrogen Energy Network (H2NET), but has drawn most heavily on the following studies:
(i) Lakeman and Browning (2001): Global Status of Hydrogen Research1 carried out by DERA for
the DTI and which includes a state of the art overview of hydrogen energy technologies,
(ii) Padro and Putsche (2001): Survey of the economics of hydrogen technologies2 carried out by
NREL for the US Department of Energy and which includes an attempt to produced levelised cost
comparisons for the major production and distribution technologies,
(iii) Ogden (1999): Prospects for building a hydrogen energy infrastructure3 which considers a wide
range of possible architectures for hydrogen energy systems.
US $/GJ has been adopted as the standard measure for comparing cost estimates, since this is the approach
adopted in (ii) above (at current exchange rates, this is approximately a factor of 4 greater than the
equivalent cost in p/kWh). Comparisons of common energy equivalents are presented in Appendix 1.

Page 3
Hydrogen
Hydrogen is an energy vector, not a primary fuel. It does not exist in pure molecular form naturally on
Earth (although hydrogen atoms are estimated to comprise 0.14% by weight of the earth's crust), but must
be produced from other sources.

Hydrogen has more energy per unit mass than any other fuel and it avoids or substantially reduces CO2 and
other emissions at the point of use (which makes it attractive for transport applications). Hydrogen can be
produced, stored, and used in many diverse ways, but the ultimate energy paths can be complex and the
most commonly suggested fuel cycles (usually involving electrolysis) are expensive, by comparison with
existing fossil fuel prices. It is interesting that some of the less talked about production routes (from fossil
fuels) are relatively inexpensive and the perceived environmental problems (continued need for extensive
carbon dioxide sequestration) may be soluble. The basic properties of hydrogen are tabulated in Appendix
2 and compared with the equivalent data for petrol and natural gas.

Hydrogen is safer than commonly held perceptions might suggest. Leaks are difficult to detect because the
gas is colourless and odourless, but it mixes much faster with air than either methane or petrol vapour (see
high diffusion coefficient) which makes accidents in the open air less critical. In confined spaces, there are
potential problems, which should be addressed by the provision of suitable vents. Hydrogen has a much
wider explosive range in air (13%-59%) than methane (6.3%-14%), but the latter is explosive at lower
concentrations. There is a factor 10 difference in minimum ignition energy between hydrogen and either
methane or petrol, but this is of little significance when it is considered that even the spark from a static
electric discharge has sufficient energy to ignite natural gas. In all other respects, hydrogen is broadly
similar to the fuels it might replace.

Hydrogen's main drawback is that of low density and consequent low volumetric heat content
(approximately 25% that of natural gas).

The principal drivers behind the current interest in the hydrogen economy are:
- oil and gas resource depletion,
- global warming,
- urban air quality,
- security of energy supply,
- lack of suitable large scale electricity storage media.

Page 4
Hydrogen production
Currently hydrogen is used almost exclusively as an industrial chemical, within which capacity it is applied
to a wide variety of uses, including ammonia production (for fertiliser manufacture), refinery use for
desulphurisation and other processes, and methanol production. The annual world production is around 500
billion Nm3. The largest producer is Air Products, who operate more than 50 individual plants, producing
over 25 million Nm3 per day, and 7 pipeline systems totalling more than 340 miles1.

The bulk of hydrogen (almost 50%) is produced by steam methane reforming (SMR), which is the most
economical (large scale) route. Partial oxidation of hydrocarbon fuels can be competitive where a cheap
source of oxygen is available.

The US DOE hydrogen production cost target2 from fossil fuel based sources is 5.70-7.60 $/GJ.

Steam methane reforming (SMR)

Four stage process:


(i) feedstock preheating and purification (needed because catalyst is highly sensitive to impurities,
e.g. sulphur, mercury, and other metals),
(ii) steam reformer
(iii) CO shift
(iv) PSA purification (absorb compounds other than H2; hydrogen recovery 80%-90%)

Reformer reaction (for methane):

CH 4 + 2 H 2 O ↔ CO2 + 4 H 2 ( ∆H 0 = +164kJ / mol ; typical operating temperature 850 oC)


CH 4 + H 2 O ↔ CO + 3H 2 ( ∆H = +205kJ / mol )
0

- operates at < 40 bar (Jacobs plant typically 20-30 bar)


- highly endothermic reaction
- conversion favoured by higher temperature and steam; conversion reduced by higher pressure
- need highly active nickel catalysts
- possibility to enhance reaction by adsorption of CO2, enabling reaction temperature to be reduced to
550 oC

CO shift reaction (favoured by lower temperature):

CO + H 2 O ↔ CO2 + H 2

- uses CO shift catalysts: iron oxide based (conventional HT 340-460 oC), (MT) iron + copper oxide
(modified HT 310-370 oC), copper, zinc, aluminium (low T 180-280 oC)
- small and medium-sized plants have single MT shift reactor
- large plants have 2 x MT reactors or HT plus MT reactors

Typical plant sizes:

Small 500 - 3000 Nm3/hr


Medium up to 25,000 Nm3/hr

1
Hydrogen Production, Supply, and Distribution, C. Spilsbury, Air Products plc, H2NET Hydrogen Production meeting,
University of Glamorgan, 14 February 2001
2
A multi year plan for the Hydrogen R&D Programme rationale, Structure and Technology Roadmaps, Office of Power
Delivery, Office of Power Technologies Energy, Energy Efficiency, and Renewable Energy, US DOE, August 1999 - cited
in Lakeman and Browning (2001)1

Page 5
Large over 25,000 Nm3/hr
Very large over 150,000 Nm3/hr
World max. 300,000 Nm3/hr (tar sands project)
Potential max. 950,000 Nm3/hr

Partial oxidation

Hydrogen may be formed by the non-catalytic partial oxidation of hydrocarbons. Any hydrocarbon
feedstock that can be compressed or pumped may be used. Overall process efficiency is only 50% (c.f.
SMR at 65%-75%)3. Pure oxygen is required as a feed.

Partial oxidation - Reformer reaction:


1
Natural gas : CH 4 + O2 → CO + 2 H 2 (1350 oC) - Syngas (Synthesis Gas) process
2
1
Coal : C + O2 → CO (1350 oC)
2
Partial oxidation - CO shift conversion:
CO + H 2 O ↔ CO2 + H 2

- uses any fossil fuel and can operate at high pressure (> 100 bar)
- BP Amoco Bulwer Island Refinery (Australia) POX unit produces 35,000 Nm3/hr, which is very low
for a POX unit (not normally economic)

A fuel processor is a mechanical device that uses heat and a catalyst to initiate changes in the chemical
composition of a hydrocarbon to release hydrogen molecules.

HydrogenSource4 lists three technologies:


(i) catalytic steam reforming (CSR) involves reacting a hydrocarbon fuel and steam in the presence
of a catalyst - an external heat source is required, the process is high efficiency,
(ii) auto thermal reforming (ATR) involves reacting a hydrocarbon fuel and steam in the presence of
a catalyst and oxygen - some of the fuel is used to generate the required heat for the reaction, the
process can be used on many different types of fuel,
(iii) catalytic partial oxidation reforming (CPOX) is similar to auto thermal reforming but allows
use of a simpler and smaller operating system.

ICI have developed a gas-heated reforming5 (GHR) process to reduce the size of the large steam system
required when methane steam reforming.

Note that the oil majors have a vested interest in making sure that the reformer is placed on the vehicle
(they can then deliver a liquid fuel with relatively little change to their existing infrastructure), which
makes CO2 sequestration much harder to achieve. General Motors recently announced they had developed
such an on-board gasoline reformer. However, there is considerable difficulty in ensuring that when the
driver wants to accelerate the reformer-fuel cell system can deliver the required response.

Methanol is easier to reform (and the process can be carried out at lower temperatures) than either kerosene
or diesel.

Technology developments are concentrating on:

3
Vezirogulu and Babir (1998), Kirk-Othmer (1991) and Leiby (1994) quoted in Padro and Putsche (1999)2
4
HydrogenSource is an independent company formed between International Fuel Cells and Shell Oil Company (see
http://www.hydrogensource.com/whoweare.html)
5
See Synetix Web site : http://www.synetix.com/methanol/technology-ghr.htm

Page 6
(i) reactor design (e.g. use of micro-channels to enhance heat and mass transfer capabilities currently
being developed for small scale reformers, e.g. Pacific Northwest National Laboratory 1 kW steam
reformer),
(ii) plasma reformation (MIT) can be applied to POX, SR, and pyrolysis,
(iii) compact design to incorporate reforming and water gas shift operations in single stage,
(iv) outlet gas clean-up (required for PEMFC operation).

Integrated gasification combined cycle (IGCC)

In an IGCC system, a coal gasifier converts pulverised coal into a synthesis gas (mixture of H2 and CO) by
adding steam and oxygen. This Syngas is then cleaned of impurities and used to produce energy in a gas
turbine. (Alternatively the product gas could be used to produce hydrogen, chemicals, or other fuels.)
Waste heat from the gas turbine is used in a steam turbine to generate more electricity. Integrating gasifier
technology with a combined cycle in this way offers high system efficiencies and ultra-low pollution levels.
Systems are designed to handle a variety of feedstocks, including low and high sulphur coal, anthracite, and
biomass. Typical systems range in size from 200-800 MWe; although modular designs of 50-150 MWe are
expected to be the basis for future plants6. Typical plants offer an improvement of 10% in thermal
efficiency over conventional coal-fired stations. The operating efficiency is between 29-41%, depending on
the fuel characteristics (ie sulphur content, ash content and calorific value), type of IGCC system (ie
entrained, moving-bed or fluidised-bed) and peak gas turbine temperature. Due to high costs and low
efficiencies, IGCC remains a demonstration technology, but it is hoped that second-generation technologies
will realise efficiencies from 45-50% and reduced costs.

Capture of CO2 is considered difficult without incurring significant efficiency and cost penalties7. Possible
future technological developments include Entrained Flow Gasifiers and Water-wall Cooled Gasifiers.

Pyrolysis

Hydrocarbons can be converted to hydrogen without producing carbon dioxide, if they are decomposed at a
sufficiently high temperature in the absence of oxygen (pyrolysis). For example methane can be "cracked"
in the presence of a catalyst such as carbon felt (or novel carbon families, such as C60 soots, graphite, or
activated carbons). Carbon black can either be sequestered or used further by a number of industries (e.g.
car tire production, metallurgic industries). This process has been developed commercially by Kvaerner
under the process name "Kvaerner Carbon Black and Hydrogen Process" (KCB&H). A pilot plant has been
operating since 1992 and a commercial plant (first stage produces 50 million Nm3/year of hydrogen and
20,000 tonnes/year of carbon black) started operation in Canada during mid-1999. The process is projected4
to be slightly more expensive than steam methane reforming before carbon black revenues are taken into
account; with these revenues it may even be cheaper. Cost projections have been presented for plants
producing 100 million Nm3/year and 500 million Nm3/year of hydrogen.

In principle, pyrolysis could also be applied to more complex hydrocarbons, biomass, and municipal solid
waste.

6
See IEA GREENTIE : http://www.greentie.org/class/ixa10.htm
7
John Griffiths, Jacobs Consultancy, IChemE Sourcing the Hydrogen Economy meeting, October 2001

Page 7
Table 1 : Typical plant sizes for hydrogen production

Type of plant Typical plant size Future developments


SMR - small 500-3000 Nm3 /hr
SMR - very large < 150,000 Nm3 /hr
POX - reactor design
- plasma reformation
- compactness
- outlet gas clean-up
Integrated Gasification Combined Currently 200-800 MWe
Cycle (IGCC) Modular designs of 50-150 MWe
are expected in near future
10% more efficient than
conventional coal stations
Pyrolysis Modular design
5,000 Nm3/hr (demonstration)
100-40,000 Nm3/hr (cost-
estimated)
Water electrolysis Modular units

Page 8
Table 2 : Reference costs for hydrogen production from more conventional technologies
(derived from Padro and Putsche, 19992)

Facility size References Specific TCI Hydrogen


(106 Nm3/day) ($/GJ) price ($/GJ)
Steam methane reforming − 48% of world hydrogen production (1998)
0.27 (small) Leiby 1994 in Padro and Putsche (1999)2 27.46 11.22
1.34 (large) Leiby 1994 in Padro and Putsche (1999)2 14.74 7.46
2.14 Leiby 1994 in Padro and Putsche (1999)2 12.61 6.90
2.80 Kirk-Othmer 1991 in Padro and Putsche (1999)2 9.01 6.26
6.75 Foster-Wheeler 1996 in Padro and Putsche (1999)2 10.00 5.44
25.4 Blok et al 1997 in Padro and Putsche (1999)2 10.82 5.97

Coal gasification - economic where oil and/or natural gas is expensive


2.80 Kirk-Othmer 1991 in Padro and Putsche (1999)2 34.2 11.57
6.78 Foster-Wheeler 1996 in Padro and Putsche (1999)2 33.1 9.87

Partial oxidation of hydrocarbons


0.27 (small) Feedstock - coker off-gas (Leiby 1994 in Padro and 21.96 10.73
Putsche (1999)2)
1.34 (large) Feedstock - coker off-gas (Leiby 1994 in Padro and 11.24 7.39
Putsche (1999)2)
2.14 Feedstock - coker off-gas (Leiby 1994 in Padro and 9.63 6.94
Putsche (1999)2)
2.80 Feedstock - residual oil (Kirk-Othmer 1991 in 22.28 9.83
Padro and Putsche (1999)2)

Biomass gasification - costs vary depending on gasifier technology


0.720 Mann 1995 in Padro and Putsche (1999)2 38.19 13.099
2.16 Larson 1992 in Padro and Putsche (1999)2 20.60 8.69
2.26 Larson 1992 in Padro and Putsche (1999)2 26.91 10.03

Biomass pyrolysis
0.024 Mann 1995a in Padro and Putsche (1999)2 30.66 12.73
0.243 Mann 1995a in Padro and Putsche (1999)2 19.31 10.11
0.811 Mann 1995a in Padro and Putsche (1999)2 16.74 8.86

Methane pyrolysis - Gaudernack and Lynum (1998)4


SMR without CO2 sequestration 6.0
SMR with CO2 sequestration 7.5
CB&H process without carbon black revenue 10.6
CB&H process with carbon black revenue 5.8

8
More expensive than coker off-gas due to increased equipment for feed handling and impurity removal
9
Delivered hydrogen price depends on biomass feedstock costs which can vary widely from expected price for dedicated
biomass production ($46.30/dry tonne) to the price for waste product ($16.50/dry tonne)

Page 9
Water electrolysis

Hydrogen can be produced from water by electrolysis. If the electricity is produced from a renewable
technology (such as solar, hydro, wind, or tidal) then the process is carbon-free. The electrochemical
splitting of water is well known:
1
H 2 O → H 2 + O2
2
Commercial electrolysis plants typically achieve efficiencies of 70-75%.

There are two basic types of electrolyser:


(i) liquid alkaline (easy to scale up, easier thermal management by circulation of electrolyte, likely to
remain popular choice for large plants in foreseeable future),
(ii) proton exchange membrane (PEM) (driven by small to medium-sized applications, such as home-
based systems, single vehicle refuelling).

Typical operating pressures are up to 50 bar (750 psig) which is insufficient for charging high pressure
cylinders.

Electricity consumption of the electrolysis process can be reduced in principle by operating at high
temperatures (900 - 1000 oC). A high temperature steam electrolyser (HTSE) is currently under
development at the Lawrence Livermore National Laboratory.

In order to achieve high energy storage densities it is necessary to store hydrogen (and oxygen for closed
regenerative systems) at high pressure. This can be achieved through the use of compressors or high
pressure electrolysers (which requires reactor containment in a pressure vessel, see current US research).

Reversible fuel cell

Attractive in principle, but limited so far in realisation due to the fact that electrode design (particularly
with regard to catalysts) is different for the electrolysis and fuel cell energy flows.

Hydrogen bromide electrolysis

The electrolysis of hydrogen bromide to hydrogen and bromine:


2 HBr → H 2 + Br2
requires half the voltage of the water electrolysis reaction and is therefore potentially more efficient. The
reaction is reversible and can therefore be incorporated in a closed loop energy storage/generation system.

This concept can be thought of as an adaptation of the RegenesysTM regenerative fuel cell being developed
by Innogy Technology Ventures Ltd in the UK, which uses a bromide/bromine cathode (i.e. one half of the
proposed system), by the addition of a hydrogen anode. This technology is being investigated by ITVL in
collaboration with SRT in the US under US DOE funding.

A further potentially useful development of this technology is that hydrogen bromide can be produced by
the reaction of methane with bromine:
CH 4 + 2 Br2 → C + 4 HBr
with the production of carbon black. The bromine can be recycled after the hydrogen bromide electrolysis
reaction.

Page 10
Hydrogen production during manufacture of chlorine

Hydrogen is produced as a by-product in the manufacture of chlorine by electrolysis of brine in a mercury


based membrane cell. Chlorine and hydrogen are explosive and so must be kept apart when evolved.
2 NaCl → 2 Na + Cl 2
2 Na + 2 H 2 O → 2 NaOH + H 2
Brine is introduced to the anode side of the electrolysis cell and the chlorine ions are oxidised to produce
chlorine gas; the sodium ions pass through the central membrane and combine with hydroxide ions to
produce sodium hydroxide (caustic soda); while the hydrogen ions are reduced to give off hydrogen gas.
The process takes place at 75-90 oC.

Global production of chlorine is 40.0 Mt/year (producing approximately 1.0 Mt/year of H2) and is growing
2%-3% per annum. In the UK, production is about 1.06 Mt/year (with 0.03 Mt/year of H2). 35% of the H2
is used as a chemical feedstock, 50% as fuel, and the remaining 15% is currently unused (vented)

The H2 gas is saturated and contains NaOH and mercury impurities.

Photoelectrolysis

Photoelectric systems split water directly into hydrogen and oxygen using sunlight as the only energy input.
Such a system does away with the need for external wiring, convertors, etc. The system must generate
sufficient internal voltage to decompose the water molecules. Multi-junction cell technology, developed by
the photovoltaics industry, involves the series connection of photovoltaic layers with different
semiconductor band gaps to produce a single cascade device. This configuration uses a higher proportion of
the solar spectrum than a simple photovoltaic cell and therefore has a higher theoretical conversion
efficiency. The cascade structure minimises surface area requirements. Amorphous silicon (a-Si) is being
considered for this technology, but stability in an aqueous environment is a key research area. See work at
NREL, Golden, Co., and Hawaii Natural Energy Institute.

DOE hydrogen production cost target10 9.50-14.20 $/GJ (US $10-15 per million BTUs).

DOE photo-conversion efficiency target 15%.

Biological hydrogen production

Hydrogen can be produced biologically in two distinct ways:


(i) photosynthesis process,
(ii) fermentation process.

Green algae capture energy from sunlight. Under anaerobic conditions, the green algae produce a
hydrogenase enzyme which produces hydrogen from water by a process known as bio-photolysis. The
conditions must be carefully regulated, since the hydrogenase enzyme works in the dark phase and is
sensitive to the presence of oxygen produced from photosynthesis. A two stage process is used to maximise
hydrogen yield. The major research challenges are:
(i) increase hydrogen production by a factor of 10, or better,
(ii) increase solar energy conversion efficiency from 5% to at least 10%,
(iii) producing a non-living membrane cell with oxygen- and hydrogen- producing enzymes.
(Hydrogen can also be produced photosynthetically by harnessing nitrogenase reactions on blue-green
algae and photosynthetic bacteria, but the reaction is inhibited by nitrogen, ammonia, and oxygen, and the

10
A multi year plan for the Hydrogen R&D Programme rationale, Structure and Technology Roadmaps, Office of Power
Delivery, Office of Power Technologies Energy, Energy Efficiency, and Renewable Energy, US DOE, August 1999 - cited
in Lakeman and Browning (2001)1

Page 11
production rate is 1000 times slower than the hydrogenase process.) Research work is progressing under
IEA Task 15 (Japan, Norway, Sweden, US, Canada).

The second main biological pathway to hydrogen uses fermentation without need for light. This is a dark,
anaerobic process carried out by many species of bacteria, one of which, Clostridia, has been singled out
for particular attention. The reaction involves a hydrogenase enzyme acting on carbohydrates to produce
hydrogen (and carbon dioxide):
C6 H 12O6 + 2 H 2 O → 2CH 3COOH + 2CO2 + 4 H 2
Laboratory studies have concentrated on pure microbial cultures, sterile feedstocks, and batch operation.
Commercialisation requires use of stable, mixed micro-flora, non-sterile feedstocks, and continuous
process operation. Potential feedstocks include sugar cane or the organic fraction of muncipal waste. The
theoretical yield is 0.5 m3 H2/kg carbohydrate, although this is yet to be achieved, largely because the
reaction is hindered by increasing hydrogen concentration. University of Glamorgan researchers have
achieved 36% of theoretical maximum in a continuously operating reactor using mixed micro flora and
non-sterile feedstocks by sparging the reactor with inert gas. Fermentative bacteria can multiply rapidly
and produce high quantities of hydrogen, but the design and operational parameters are not yet well
established.

Page 12
Table 3 : Reference costs for hydrogen production from electrolysis and less conventional
technologies (derived from Padro and Putsche, 19992)

Facility size References Specific TCI Hydrogen


(106 Nm3/day) ($/GJ) price ($/GJ)
Electrolysis- non-specific technologies
0.096 Andreassen 1998 in 2 31.88 28.7
2.8 Kirk-Othmer 1991 in 2 2.95 (?) 20.6
6.75 Foster-Wheeler in 2 30.97 24.5

Electrolysis - wind based


0.247 (0.02311) Mann et al 1998 in 2 − technology 200012 158.6 (165.7) 20.2 (22.6)
0.279 (0.028) Mann et al 1998 in 2 − technology 2010 92.5 (91.1) 11.0 (12.6)

Electrolysis - solar photovoltaics based


0.195 (?) Mann et al 1998 in 2 − technology 2000 485.8 41.8
0.209 (?) Mann et al 1998 in 2 − technology 2010 242.0 24.8

Mann et al. (1998)5 observe that economics of renewable powered electrolysis can be improved by:
- use of off-peak electricity when wind/solar energy not available,
- sale of electricity from wind/solar energy during peak tarif times (i.e. suspend hydrogen
production).
For the US cases studied, a combination of these can potentially improve economics by a factor of 2
(dependent on wind/solar regime, electricity tarifs, feed-in contracts, etc.). It is desirable to evaluate case
studies for equivalent situations within the UK electricity network and this will be carried out as part of the
Phase II study.

11
Author's spreadsheet calculation based on data given in original paper - the original capacity throughput is incompatible
with the size of the wind farm cited (10 MW)
12
Mann et al note the potential to increase revenue from these plants by utilising a grid connection to purchase power at off-
peak periods when the renewable generators were not operating and/or to sell electrical power at peak time (at the cost of
reduced hydrogen production)

Page 13
Hydrogen storage and distribution
To achieve a fully functioning hydrogen economy, hydrogen will need to be stored on a wide range of
scales. Large, centralised storage would be required if hydrogen is produced in large plants for wider
distribution; longer term or seasonal storage would be required in systems relying on large penetrations of
renewable energy; comparatively small scale storage is required on board vehicles, possibly in homes, and
for portable devices.

Technology Specific developments Targets


Automotive applications 6.5 wt% (US DOE)
62 kg H2 m-3
Portable power systems (military) 11.0 wt%
Stationary systems likely to be cost driven

Surface storage in large


compressed gas tanks
Storage in depleted gas wells, salt Already used for natural gas
caverns, etc storage
Novel composite cylinders
Liquefied hydrogen
Conventional metal hydrides Ambient temperature hydrides Little improvement in capacity
FeTi, LaNi5, typically < 2.0 wt% likely
- improve cycle life
- improve impurity tolerance
- improve kinetics
- improve T of operation
UBirm using palladium,
platinum, ruthenium to improve
kinetics and impurity tolerance
Higher temperature hydrides
usually contain Mg (e.g. Mg2Ni),
can store up to 7.0 wt%
Novel metal hydrides Sodium aluminium hydrides with 5.0 wt% (IEA Task 12)
titanium or zirconium catalyst
(e.g. NaAlH4, Na3AlH6)
- titanium doping makes reaction
reversible
- currently slow kinetics
Lithium beryllium hydride,
reversible hydrogen capacity of
almost 9.0 wt%, but requires T >
250 oC to achieve useful
pressures
Carbon nano-adsorbents (nano- Claims of 0.0 to > 50.0 wt% have Work continuing under IEA
fibres, nano-tubes, etc) been made - a growing consensus Annex 12
appears to be that early results
were due to water absorption and
only a few wt% will be achieved,
in practice
Hydride hydrolysis
Glass microspheres

Page 14
Compressed gas

Conventional steel cylinders are heavy and bulky. In a typical system, hydrogen is compressed to about
20.7 MPa (i.e. 200 bar / 50 litre), requiring around 2.3 kWh/kg13. Larger, spherical pressure vessels are
used where larger volumes of hydrogen are required (say > 14,000 Nm3).

Composite cylinder technology - inner aluminium or thermoplastic liner providing gas barrier, strengthened
by overwrapping layers of glass or carbon fibre reinforcement, lighter weight. Initial cylinders
manufactured by EDO and Luxfur realised around 3.0 wt% H2 at pressures 200-300 bar.

Lawrence Livermore National Laboratory (with IMPCO Technologies, Thiokol Propulsion, and Aero Tec
laboratories) have produced a prototype composite cylinder which can store 11.3 wt.% H2 at a working
pressure of 5000 psi [345 bar] (equating to 3.8 kWh/kg specific energy storage). Burst pressure safety
margin was found to be 2.15 (2.25 design). Development plans include a 10,000 psi unit (but note higher
compression cost and complexity) and examining way to deal with heating problems during fast filling of
the cylinders. There is a trade-off between improved gravimetric performance from using higher grade
carbon fibre and the overall cost per tank. Preliminary cost estimates suggest that at high volume
production (>500,000 units per year) tanks holding 3.6 kg H2 at 345 bar will cost $640 (£444), the largest
cost component being high grade carbon fibre.

The requirement for additional compression consumes energy, adds to expense, and has a significant
footprint. Compressor development is required in the areas of new materials, improved efficiency, and
more compact designs.

Problems also arise with heating of tanks on fast filling, which may limit the rate at which tanks can be
refilled. Pre-cooling of the hydrogen or the use of heat exchangers may be required to overcome this
barrier.

Thiokol Propulsion is also considering designs for conformable tanks to fit existing vehicle envelopes,
since it is often difficult to accommodate cylindrical tanks. This may not be such a problem if the vehicle is
designed around the tank.

Liquefaction

The energy required for liquefaction14 ranges between 8.5 kWh/kg and 13.0 kWh/kg, depending on plant
size. Energy requirements as low as 5.0 kWh/kg may be achievable in future magnetocaloric cooling15.
Very efficient tank insulation (cryogenic dewar vessels) is required to keep temperatures below the boiling
point of hydrogen (only 20 K at 1 bar). Spherical shape reduces surface area to volume ratio and hence heat
losses, but are difficult to accommodate in a car envelope. Insulation increases volumetric size of tanks.

Sizes range from several litres to 3,800 m3 (the largest in the world at NASA).

Transport and stationary infrastructures can potentially be linked.

The value of intermittent renewable power sources can be enhanced by use of hydrogen as an energy
vector.

Boil-off rates quoted by BMW are around 3% per day, but this is only a problem if the hydrogen is stored
for longer periods of time, since in normal applications the gaseous phase will be used first.

13
Wurster, R., Zittel, W., http://www.hydrogen.org/Knowledge/Ecn-h2a.html, section 9.7
14
This includes the energy required to convert all the -ortho form of H2 to the -para form (at ambient temperature, the gas is
75% -ortho and 25% -para)
15
Amos, W.A., 1998, p. 23, and Wurster, R., Zittel, W., http://www.hydrogen.org/Knowledge/Ecn-h2a.html, section 9.8

Page 15
An alternative to cryogenic storage at ambient pressure is to use insulated pressure vessels with the
capacity to operate at LH2 temperatures and high pressure (24.8 MPa; 3600 psig). These vessels can accept
either LH2 or CH2 as fuel. Capacity, and hence range, are reduced to one-third of the LH2 capacity when
using CH2. The cost of liquefaction need only be paid where journey ranges in excess of 200 km are
required.

Solid state hydrogen storage

• metal hydride
• intermetallic compounds (LaNi5, ZrMn2, TiFe) - can achieve around 2.0 wt/% H2; to achieve
higher need to use other materials
• solid solution alloys (V-Ti-Fe, V-Ti-Mn)
• magnesium (nanocrystalline) - high T (~300 oC), but quite slow kinetics (even after surface
treatment), nanotechnology with low T and p ~7.0 wt.% seems possible (check)
• carbon-based
• cryo-adsorption (77 K, 60 bar)
• nanotubes, nanofibres (C60 required high T, so researchers looked at other structures, 10.0
wt.% to 50.0 wt.% originally claimed, but not independently verified and appear to have been
due to water absorption) - mechanism has yet to be proved, but thought to be due to hydrogen
condensing in the spaces between the graphite layers of the fibre
• novel hydride
• hydride complexes (Mg2NiH4, NaAlH4)
• non-reversible (sodium borohydride with product recycling) - used by Millennium Cell and
Powerball
NaBH 4 + 4 H 2 O → NaOH + H 3 BO3 + 4 H 2

Metal hydride storage systems

Two classes :
(i) those which operate at ambient temperature,
(ii) those which require higher temperatures (> 200 oC)

Target : 7 wt.% of H2
Low desorption temperature (< 100 oC)
Low operating pressure
Ability to recycle many times
Fast recharge kinetics

Hydride hydrolysis

Hydrogen can be liberated by the action of water on certain hydride materials:


MH + H 2 O → MOH + H 2
Historically water has been added to solid powder hydrides, giving rise to an exothermic reaction which is
difficult to control. Current research is concentrating on fluid forms of the hydride, which improves heat
and reaction control. There are two approaches:
(i) stabilisation of the hydride in an aqueous hydroxide solution (Millennium Cell LLC using sodium
borohydride and Pt or Ru catalyst; the reaction product sodium metaborate is a safe compound
found in laundry detergents and can be recycled at centralised processing plants; theoretical
storage of 10.8 wt%, values of 7.0 wt% claimed),

Page 16
(ii) stabilisation of the hydride in an organic oil slurry (Thermo Technologies, USA using light metal
hydride such as lithium hydride in an organic dispersant such as light mineral oil; likely efficiency
of between 6.4 wt% and 15.3 wt%).

Glass microspheres

Hydrogen can be contained at very high pressures in small glass microspheres, which perform as tiny
pressure vessels. Research funded by the US DOE has recently been suspended following a detailed cost
and performance analysis.

Relative merits of hydrogen storage systems and comparison of costs

The two main factors affecting the cost of a hydrogen storage system are production rate and storage
time. The required production rate determines the size of compressors and liquefaction plants and their
operating costs (electricity, heating/cooling requirements); the production rate multiplied by the number of
storage-days gives the overall capacity, which in turn determines the unit size and capital cost.

Amos (1998) 6 compares the relative merits and costs of competing hydrogen storage technologies. When
considering final hydrogen delivered cost it is necessary to take account of delivery distance, as well as
production cost, production rate, and storage time, making the optimum economic storage solution case-
dependent. Nonetheless, it is possible to draw some general conclusions:

• Underground storage is the cheapest method at all production rates and storage times (due to low
capital cost of the cavern): biggest cost item is electricity cost for compression; relatively insensitive
to changes in production rate and storage time; additional transport cost to consumer may be high,
but underground storage may have applications for seasonal storage or security of supply.

• Metal hydride storage is perceived to have no economy of scale (high capital cost of storage alloy),
so does not compete with other options at high production rates or long storage times, but may be
ideal at low flow rates and short storage times. Since it is also considered the safest storage option,
this makes it a leading candidate for on-vehicle storage, subject to achieving satisfactory energy
densities.

• Liquid hydrogen storage is not economic at low production rates (high capital cost of liquefier) and
struggles to compete with compressed gas at higher production rates unless longer storage times are
required, when the lower capital cost of liquid hydrogen dewars compared to compressed gas
pressure vessels becomes the chief factor. The comparatively low dewar capital cost makes liquid
hydrogen storage costs relatively insensitive to storage time. At very high production rates, economy
of scale reduce storage costs until they are limited by the electricity costs associated with
liquefaction.

• Compressed gas hydrogen storage is competitive with liquid hydrogen and metal hydride storage
for small quantities and low production rates. At low production rates, the capital cost of the
pressure vessel dominates, at higher volumes the critical factor is electricity cost for compression.
As storage time increases, the capital cost of the pressure vessel begins to dominate the cost. A
further option is to increase the operating pressure of the system (smaller, lower cost tank; higher
compressor capital and compression running costs): for short times, there is a balance between these
costs, at longer times the capital cost reduction is the dominant factor resulting in an optimum at
maximum operating pressure.

Page 17
Table 4 : Reference costs for hydrogen storage systems (derived from Padro and Putsche,
19992)

Facility size References Specific TCI Hydrogen


(GJ) ($/GJ) price ($/GJ)
Compressed gas storage − short term (1-3 days)
131 9,008 4.21
13,100 2,992 1.99
20,300 2,285 1.84
130,600 1,726 1.53
Compressed gas storage − long term (30 days)
3,900 3,235 36.93
391,900 1,028 12.34
3.9 million 580 7.35

Liquefied hydrogen − short term (1-3 days)


131 35,649 17.12
13,100 7,200 6.68
20,300 1,827 5.13
130,600 3,235 5.26
Liquefied hydrogen − long term (30 days)
3,900 1,687 22.81
108,000 1,055 25.34
391,900 363 8.09
3.9 million 169 5.93

Metal hydride − short term (1-3 days)


131 4,191 2.89
130,600 18,372 7.46
Metal hydride − long term (30 days)
3,900 18,372 205.31
3.9 million 18,372 205.31

Underground hydrogen storage − short term (1 day)


1.0-5.0

Page 18
Hydrogen distribution and transport
Hydrogen is conventionally distributed in gaseous form in cylinders or large tanks, or in liquid form by
tanker or occasionally pipeline. Future applications may include storage and transport as a solid metal
hydride. The optimum method depends on quantity delivered and distance transported.

A large scale hydrogen pipeline distribution infrastructure is conceivable, but would be expensive. There
are significant technical problems (arising from embrittlement of metals and the use of certain kinds of
plastic pipe) related to the use of the existing natural gas pipeline network (although it is considered
feasible to dilute the natural gas supply by up to 10% by volume with hydrogen). However, a complete
changeover to pure hydrogen distribution is probably logistically impractical since some consumers would
still require natural gas during any changeover period.

Some researchers have suggested that local production of hydrogen − by steam methane reforming or
electrolysis of water − may be economically feasible.

Table 5 : Hydrogen transmission and transportation costs (derived from Padro and Putsche,
19992)
Transmission/ References Specific TCI Hydrogen
Transport ($/GJ) price ($/GJ)
distance (km)

Hydrogen transmission costs (pipeline) − transmission rate 0.15 GW (2 studies)


161 14.14/21.22 2.03/2.83
805 67.53/106.24 8.87/13.84
1609 134.18/210.32 17.41/27.23

Hydrogen transmission costs (pipeline) − transmission rate 1.5 GW (47.3 million GJ/year) (2 studies)
161 2.13/2.83 0.49/0.83
805 7.47/11.59 1.17/2.09
1609 14.13/22.3 2.03/3.53

Hydrogen transport as liquefied hydrogen by truck − 45,418 - 45.6 million GJ/year


16 0.44-11.0 0.24-1.60
161 0.77-11.0 0.52-1.84
805 2.70-11.0 2.00-3.10
1609 5.10-11.0 3.90-4.70

Hydrogen transport as compressed gas by truck − truck tube unit capacity 21.72 GJ
16 4.10 4.70
161 8.20 10.60
805 30.20 41.10
1609 57.60 79.40

Hydrogen transport via truck using metal hydrides − metal hydride at $18,375/GJ
16 7.54 2.63
161 15.08 5.75
805 55.28 21.92
1609 105.54 42.11

Page 19
Hydrogen end use systems
Car (fuel cell)7 : 0.037 GJ/day (3.4 Nm3/day)
Bus (fuel cell)7 : 2.4 GJ/day (222 Nm3/day)

Hydrogen-fuelled internal combustion engines

Hydrogen is a good fuel for IC engines and can improve efficiency by around 20% compared to the use of
gasoline. Although still limited by the Carnot efficiency, this is comparable with the efficiencies achieved
in practice with PEM fuel cells.

The ideal thermal efficiency of an IC engine is given by:


k −1
1
η =1 −  
r
where r = compression ratio and k = ratio of specific heats (Cp/ Cv). In hydrogen engines both ratios are
higher due to hydrogen's lower self-ignition temperature and ability to burn in lean mixtures. The downside
in a standard engine is a loss of power due to the lower energy content of the fuel/air mixture in the
combustion chamber. Although NOx emissions are produced, they can be limited to an order of magnitude
less than from a petrol engine.

Work is needed to develop more advanced fuel injection techniques and to re-design combustion chamber
and cooling system topography to hydrogen characteristics, rather than simply modifying conventional
gasoline engines.

Hydrogen fuelled turbines

Hydrogen use in turbines and jet engines is similar to use of conventional jet fuel. There are some avoided
problems from sediment and corrosion and considerably lower pollutant levels. Overall efficiency
improvements are likely due to the capacity to increase gas inlet temperatures beyond the current limit of
800 oC.

Fuel cells

A fuel cell utilises a chemical process to convert hydrogen (or a hydrogen-rich fuel) into electrical energy
and heat8. It can do so at high efficiency (not subject to limitations of the Carnot cycle), producing a non-
fluctuating, DC power output. Most fuel cells (notably PEMs) have high power densities (i.e. high power
output per unit weight, volume, or area). Different types of fuel cell are distinguished by their different
electrolytes and the different temperatures reached during operation.

There are limitations to fuel cell operation which mean that the theoretical cell voltage cannot be fully
utilised: these include the oxygen deactivation barrier at the cathode, and mass transport problems caused
by diffusion of gaseous species and water through the electrodes, electrolyte, and any catalyst layers in the
fuel cell. Only about half the energy released by the reaction:
H 2 + O2 → 2 H 2 O
can be converted into useful electrical energy. A typical fuel cell will have an area of 50 cm2 and generate a
potential difference of 0.7 V. A stack of such cells is required to produce a useful overall voltage.

Fuel cells display high efficiency across most of their output power range. This means that, whereas the
internal combustion engine has a point of maximum operating efficiency (usually at high revolutions with
large fall-off at low revolutions), the fuel cell has a very flat characteristic and is generally more efficient at

Page 20
fractions of its rated power. As a consequence, the fuel cell will show even larger advantages over an IC
engine during a drive cycle compared to continuous output tests.

The main disadvantage of fuel cells is the large number of ancillary units required, the so-called balance of
plant. This equipment, for fuel supply, water removal, metering, fuel reformation (if required), pumps,
compressors, and heat exchangers, makes a 200 kW phosphoric acid fuel cell generating system several
times the volume of a similarly rated diesel or natural gas CHP unit.

Table 6 : Fuel cell characteristics16


Type of fuel cell Electrolyte Operating Applications Drawbacks/
temperature research topics

Mobile ion Typical efficiency


Alkaline Potassium 50 - 90…(200) oC Space applications CO2 intolerant
hydroxide (problem if not
(85 wt% high T) acid electrolyte)
(35-50 wt% low T)
Electrolyte
OH- retained in matrix
(usually asbestos)

Electrocatalysts
(Ni, Ag, metal
oxides, noble
metals)

Novel AFC
designs using
hydroxyl-
conducting
polymer
electrolyte
mebrane
Proton exchange Polymeric 50 - 125 oC Vehicles and Requires CO
membrane (PEM) mobile levels < 10 ppm
H+ 40% + applications
Develop to
Low power CHP increase operating
- higher power T towards 200 oC,
density than SOFC increase efficiency,
decrease CO
Start up time 1-3 s sensitivity

Catalyst typically
Pt (being reduced)
or Pt-Ru where CO
impurities exist

High T proton
conducting
electrolytes are

16
Table data assembled from Rodney Allam presentation at IChemEng meeting and Larminie and Dicks (2000)8 and
Lakeman and Browning (2001)1

Page 21
being developed
Phosphoric acid Orthophosphoric ~ 220 oC Large numbers of SiC matrix
(PAFC) acid 200 kW CHP
H+ 37%-42% (HHV) systems in use Electro-catalyst in
anode and cathode
Still low T so is platinum black
hydrocarbons
reformed
externally
Molten carbonate Lithium/potassium ~ 650 oC Suitable for Electrolyte
(MCFC) or medium to large retained in a
sodium/potassium scale CHP, up to ceramic matrix of
carbonate mixture MW capacity LiAlO2

CO32- Potential to use Noble metal


CO and CH4 as catalysts usually
fuels with catalyst not required at this
temperature
Solid oxide Stabilised zirconia 500 - 1000 oC Suitable for all Noble metal
(SOFC) sizes of CHP catalysts usually
O2- systems, 2kW to not required at this
multi-MW temperature
- more tolerant
of impurities
than PEMFC

Potential to use
CO and CH4 as
fuels
Direct methanol Sulphuric acid or 50 - 120 oC
polymer

High temperature fuel cells were originally developed to make fuel cells more compatible with direct use of
fossil fuels, but still need an external reformer (few kW - 11 MW = biggest ever built, but still a very low
throughput!). With high temperature can do internal reforming (at the anode).

US DOE cost goals for PEM fuel cells are capital cost < $500 (£350) / kW and electricity production cost
$12.5/GJ ($0.045/kWhr); unpressurised cells should be able to produce 600 mA/cm2 at 0.6 V.

Table 7 : Economics of fuel cells (derived from Padro and Putsche, 19992)
Fuel cell References Specific TCI Electricity
type/size ($/kW) price ($/kWh)
Alkaline FC

PEMFC
10 kW (today) Barbir and Gomez 1997 in Padro and Putsche 3,000 0.25-0.30
(1999)2
10 kW (future) Barbir and Gomez 1997 in Padro and Putsche 1,150 0.09-0.095
(1999)2

Page 22
7 kW (today) Plug Power 1999 in Padro and Putsche (1999)2 8,500 (not provided)
7 kW (future) Plug Power 1999 in Padro and Putsche (1999)2 4,000 0.07-0.10

PAFC
25 kW (future) Mugerwa et al 1993 in Padro and Putsche (1999)2 2,250 0.13
200 kW (today) Garche 1998 in Padro and Putsche (1999)2 3,000 (not provided)
250 kW (future) Mugerwa et al 1993 in Padro and Putsche (1999)2 2,440 0.13

MCFC
25 kW (future) Mugerwa et al 1993 in Padro and Putsche (1999)2 1,355 0.12
250 kW (future) Mugerwa et al 1993 in Padro and Putsche (1999)2 1,740 0.11
3.25 MW Mugerwa et al 1993 in Padro and Putsche (1999)2 1,330 0.10
(future)
100 MW Mugerwa et al 1993 in Padro and Putsche (1999)2 600 0.06
(future)

SOFC

Direct methanol

Page 23
Hydrogen systems
Ogden and co-workers3 have worked extensively on possible architectures for hydrogen production and
delivery systems for transport applications. Figure 1 shows potential near-term options and Figure 2 shows
long-term options.

In the short term, Ogden considers the main options to comprise:


(a) hydrogen produced from natural gas in large, centralised reforming plant and delivered to the user
as a liquid via refuelling stations,
(b) hydrogen produced from natural gas in large, centralised reforming plant and delivered to the user
via small scale pipeline networks to refuelling stations,
(c) hydrogen derived from chemical industry waste sources and delivered to the user via small scale
pipeline networks to refuelling stations,
(d) hydrogen produced at the refuelling station via small-scale steam reforming of natural gas,
(e) hydrogen produced at the refuelling station via small-scale water electrolysis.

In the long term, Ogden expects that the options might increase to include:
(a) a broadening of the fuel base for centralised hydrogen production to include gasification of
biomass, coal, or muncipal solid waste,
(b) large scale centralised electrolysis plants powered by wind, solar, or nuclear power,
(c) centralised hydrogen production from gasification of biomass, coal, or muncipal solid waste, but
including CO2 sequestration.

There is a broad consensus that steam methane reforming (SMR) is the cheapest conventional option for
large scale hydrogen production. However, there is a tendency to dismiss this route, due to the continued
emission of CO2. Kvaerner have argued in favour of SMR coupled to CO2 sequestration in large scale
plants, or the use of pyrolysis (which appears to have been neglected in many studies) with zero CO2
process emissions. Even without CO2 sequestration, it is possible that process efficiencies may result in
overall CO2 emissions reductions from fuel cycles including the SMR process.

Given the high costs of hydrogen storage and distribution, the optimisation between large scale centralised
hydrogen production compared with small scale distributed production has yet to be resolved. Development
work is continuing to reduce the costs of small to medium scale electrolysis plants and the development of
advanced steam reformer technology as part of combined heat and power systems is seriously considered,
lending an impetus to proposals for distributed systems. However, CO2 sequestration within SMR plants is
only likely to be economically feasible in large plants, which will mitigate in favour of these plants and
associated transmission grids if governments impose harsh penalties to encourage CO2 emissions reduction.

Overall economics of hydrogen systems

Taking the cost estimates above and considering only the best and worst cases, without reference to likely
technological development, a snapshot is obtained of the current economic feasibility of hydrogen energy
systems. These costs are compared with the current cost of petrol (at the pump and pre-tax) and natural gas
supplies. The comparison must be considered in the light of overall system costs and energy efficiency
("well-to-wheels") and possible future tax regimes to mitigate against the emission of greenhouse gases.

Comparing the costs of natural gas at the wellhead (1.80-5.98 $/GJ) or petrol on the open market at
$20/barrel (3.5 $/GJ) with levelised costs for hydrogen production from SMR (5.44-11.22 $/GJ) or
renewable-electrolysis systems (20-42 $/GJ), it is clear that hydrogen is expensive. However, the
assessment is based on current fossil fuel prices and must be considered against a background of falling
costs for renewable electricity generation. The work by Mann et al (1998)5 shows that costs for hydrogen
from renewable-electrolysis systems could fall considerably (11-25 $/GJ) by 2010, due to cost
improvements for large wind systems (currently being realised) and electrolyser systems (which can be
expected to benefit from technology improvements in fuel cell design and manufacture).

Page 24
The pre-tax UK pump price for petrol (i.e. including all delivery and transportation costs) is around 8 $/GJ,
rising to 31 $/GJ when tax is included. Given short term compressed hydrogen gas storage and only local
transportation, hydrogen from renewable-electrolysis could be competitive if it was zero-rated for tax, even
today.

Table 8 : Reference costs for overall hydrogen production


Facility size References Specific TCI Fuel price
(106 Nm3/day) ($/GJ) ($/GJ)
Steam methane reforming − 48% of world hydrogen production (1998)
0.27 (small) Leiby 1994 in Padro and Putsche (1999)2 27.46 11.22
6.75 Foster-Wheeler 1996 in Padro and Putsche (1999)2 10.00 5.44
Electrolysis - wind based
0.247 (0.02317) Mann et al 1998 in Padro and Putsche (1999)2 − 158.6 (165.7) 20.2 (22.6)
technology 200018
0.279 (0.028) Mann et al 1998 in Padro and Putsche (1999)2 − 92.5 (91.1) 11.0 (12.6)
technology 2010
Electrolysis - solar photovoltaics based
0.195 (?) Mann et al 1998 in Padro and Putsche (1999)2 − 485.8 41.8
technology 2000
0.209 (?) Mann et al 1998 in Padro and Putsche (1999)2 − 242.0 24.8
technology 2010
Petrol
70p/litre Pump price (December 2001) 31
17.8p/litre Estimated pre-tax (Dec 2001) 7.9
$20/barrel 3.5
Natural gas
$1.90/MBtu Average annual wellhead price (Henry Hub 1990- 1.80
99)
$6.31/MBtu Average wellhead spot price (Henry Hub Dec 5.98
2000)

17
GD spreadsheet calculation based on data given in original paper
18
Mann et al note the potential to increase revenue from these plants by utilising a grid connection to purchase power at off-
peak periods when the renewable generators were not operating and/or to sell electrical power at peak time (at the cost of
reduced hydrogen production)

Page 25
Conclusions
While most attention is focussed on producing hydrogen from renewable power sources such as wind and
solar, current cost projections clearly favour the production of hydrogen from fossil fuel sources, most
notably from methane by the SMR process. Some improvements in overall efficiency (arising largely from
the good performance of fuel cells used on an urban driving cycle) may result in reduced carbon dioxide
emissions from this latter route, but a large reduction in emissions would require sequestration at
significant extra cost.

The costs of storage and transportation of hydrogen lead many analysts to favour the development of small
scale distributed production systems (with reduced storage requirements) close to the point of end-use of
hydrogen. However, if SMR is the chosen technology and CO2 sequestration is regarded as mandatory or,
at the very least, highly advisable, then this can be achieved more cheaply in larger, centralised plants.

With regard to carbon sequestration, a still more promising route would appear to be pyrolysis (e.g. the
Kvaerner CB&H process) by which means hydrocarbons are decomposed in the absence of oxygen.
According to Kvaerner's cost projections, the cost of hydrogen from pyrolysis is only slightly greater than
from the SMR process, even before income from carbon black sales are accounted for. Of course,
sequestration of carbon black would be a more straightforward process than for CO2.

Hydrogen production by electrolysis depends on the price of electricity and the capital cost of the
electrolysis plant. It is possible that early niche markets may arise where excess off-peak electricity is
available or the costs can be used to offset the need for grid reinforcement for large renewable energy
developments.

Technological developments are required in the field of hydrogen storage and distribution.

The large current activity in fuel cell technology is expected to result in significant cost reductions for fuel
cells. Much of the technical knowledge should be relevant to reducing costs for electrolysers as well and a
more detailed study of cost projections for renewable-electrolysis is required.

Overall, it is evident that increased use of hydrogen as an energy carrier could result in substantial
improvements in greenhouse gas emissions through:
(i) sequestration of carbon or CO2 in large, centralised plants processing fossil fuels,
(ii) efficiency improvements from using fuel cells (and possibly even IC engines) with hydrogen,
(iii) generation of hydrogen using renewable energy resources and its subsequent use as a transport
fuel.

Page 26
Figure 1 : Near term options for supplying hydrogen transportation fuel (reproduced from Ogden, 19993)

Figure 2 : Long term options for supplying hydrogen transportation fuel (reproduced from Ogden, 19993)

Page 27
Appendix 1 : Energy units
OECD uses tonnes of oil equivalent (toe) as the basic unit to compare disparate energy balances. 1 toe is
defined as 107 kilocalories (41.868 GJ), equal, within a few per cent, to the net heat content of 1 tonne of
crude oil19.

1 Mtoe = 4.1868 x 104 TJ


1 Mtoe = 3.968 x 107 MBtu
1 Mtoe = 11630 GWh = average power of 1.33 GW over 1 year20
1 Mtoe = 7.3 million barrels21

1 TWh(e) = 0.086 Mtoe (primary energy equivalent for hydro, non-thermal electric)
1 TWh(e) = (0.086 / 0.33) Mtoe (primary energy equivalent for nuclear)

1 Mtce = 1 million tonnes of coal equivalent (= 0.697 Mtoe)22


1 Mtce = 8141 GWh

Coal:
1 MW (thermal power) [MWth] = approx 1,000 kg steam/hour
1 MW (electrical power) [MWe] = approx MWth / 3
A 600 MWe coal-fired power station operating at 38% efficiency and 75% overall availability will
consume approximately:
- Bituminous coal (CV 6,000 kcal/kg NAR): 1.5 Mt/annum
- Brown coal (CV 2,250 kcal/kg NAR): 4.0 Mt/annum

Hydrogen:
Lower heating value = 10800 kJ/Nm3 = 120.0 MJ/kg = 3.0 kWh/Nm3
Upper heating value = 12770 kJ/Nm3 = 141.9 MJ/kg
Density (gaseous) = 0.09 kg/m3
Density (liquid) = 70.9 kg/m3 (-252 oC)
1 Nm3 H2 = 10,800 kJ (LHV)
106 Nm3 H2 = 10,800 GJ (LHV)
106 Nm3 H2 = 0.258 x 10-3 Mtoe
1 Mtoe = 3.877 x 103 million Nm3 H2

1 kWh = 3.6 x 10-3 GJ


1 million BTU = 1.055 GJ

1 $/GJ = 0.25 p/kWh (@ $1.44 = £1.00)

19
International Energy Agency web site: http://www.iea.org/stats/files/mtoe.htm
20
Royal Commission on Environmental Pollution preferred comparison
21
Oil Industry Conversions: http://www.nepo.go.th/ref/UNIT-OIL.html
22
World Coal Institute web site : http://www.wci-coal.com/facts_conversion.htm

Page 28
Appendix 2 : Hydrogen characteristics and safety
Hydrogen characteristics23:

Units Hydrogen Natural gas Gasoline


Lower Heating Value (MJ/kg) 120 50 44.5
Auto-ignition Temperature24 (°C) 585 540 228-501
Flame Temperature (°C) 2045 1875 2200
Limits of Flammability in Air (Vol. %) 4-75 5.3-15 1.0-7.6
Minimum Ignition Energy25 (µJ) 20 290 240
Limits of Detonation in Air26 (Vol. %) 13-65 6.3-13.5 1.1-3.3
Theoretical Explosive Energy (kg TNT/m3 gas) 2.02 7.03 44.22
Diffusion Coefficient in Air27 (cm2/s) 0.61 0.16 0.05

Standard atmosphere : 14.696 lbf in-2 = 1.01325 x 105 Pa


105 Pa = 1 bar

23
IEA Greenhouse Gas R&D Programme : http://www.ieagreen.org.uk/h2rep.htm
24
Potential for spontaneous combustion (i.e. hydrogen is marginally safer)
25
Hydrogen appears more likely to ignite, but even the spark from a static electric discharge has sufficient energy to ignite
natural gas, so the factor 10 difference has little practical significance
26
Explosive range of hydrogen is greater, but methane is explosive at much lower concentrations
27
Hydrogen mixes much faster in air than natural gas or gasoline vapour; this is an advantage in the open, but a disadvantage
in enclosed spaces (hydrogen being very much lighter than air rises quickly, as does methane to a lesser extent; petrol
fumes remain close to the ground)

Page 29
Appendix 3 : Basic format for technology summaries

Item Typical data


Technology type
Specific model
Typical plant size(s)
Efficiency
Lifetime
Capacity factor
H2 throughput (Nm3/day; GJ/year)
Capital cost / unit H2 throughput
Running costs / unit H2 throughput
Selling cost of H2
Likely innovations (next 10 years)
Possible innovations (> 10 years)
Likely selling cost of H2 (after 10 years)
Likely selling cost of H2 (after 20 years)
Likely selling cost of H2 (after 50 years)

The economics are clearly affected by the required rate of return on investment and the total lifetime of the
plant.

Page 30
References

1
Lakeman, J.B., Browning, D.J., DERA, Global Status of Hydrogen Research, ETSU
F/03/00239/REP, 2001
2
Padro, C.E.G., Putsche, V., Survey of the economics of hydrogen technologies, NREL Technical
Report NREL/TP-570-27079, 1999
3
Ogden, J.M., Prospects for building a hydrogen energy infrastructure, Ann. Rev. Energy Environ.
1999, Vol. 24, p. 227-279, 1999
4
Gaudernack, B., Lynum, S., Hydrogen from natural gas without release of CO2 to the atmosphere,
Int. J. of Hydrogen Energy, Vol. 23, No. 12, p. 1087-1093, 1998
5
Mann, M.K., Spath, P.L., Amos, W.A., Technoeconomic analysis of different options for the
production of hydrogen from sunlight, wind, and biomass, Proc. of the 1998 US DOE Hydrogen
Program Review, NREL/CP-570-25315
6
Amos, W.A., Costs of storing and transporting hydrogen, NREL/TP-570-25106, November 1998
7
Ogden, J.M., Hydrogen energy systems studies, Proc. of the 1999 US DOE Hydrogen Program
Review, NREL/CP-570-26938, 1999
8
Larminie, J., Dicks, A., Fuel cell systems explained, Wiley, 2000

Page 31
The inter-disciplinary Tyndall Centre for Climate Change Research undertakes integrated
research into the long-term consequences of climate change for society and into the
development of sustainable responses that governments, business-leaders and decision-
makers can evaluate and implement. Achieving these objectives brings together UK climate
scientists, social scientists, engineers and economists in a unique collaborative research
effort.
Research at the Tyndall Centre is organised into four research themes that collectively
contribute to all aspects of the climate change issue: Integrating Frameworks; Decarbonising
Modern Societies; Adapting to Climate Change; and Sustaining the Coastal Zone. All
thematic fields address a clear problem posed to society by climate change, and will generate
results to guide the strategic development of climate change mitigation and adaptation
policies at local, national and global scales.
The Tyndall Centre is named after the 19th century UK scientist John Tyndall, who was the
first to prove the Earth’s natural greenhouse effect and suggested that slight changes in
atmospheric composition could bring about climate variations. In addition, he was committed
to improving the quality of science education and knowledge.
The Tyndall Centre is a partnership of the following institutions:
University of East Anglia
UMIST
Southampton Oceanography Centre
University of Southampton
University of Cambridge
Centre for Ecology and Hydrology
SPRU – Science and Technology Policy Research (University of Sussex)
Institute for Transport Studies (University of Leeds)
Complex Systems Management Centre (Cranfield University)
Energy Research Unit (CLRC Rutherford Appleton Laboratory)
The Centre is core funded by the following organisations:
Natural Environmental Research Council (NERC)
Economic and Social Research Council (ESRC)
Engineering and Physical Sciences Research Council (EPSRC)
UK Government Department of Trade and Industry (DTI)

For more information, visit the Tyndall Centre Web site (www.tyndall.ac.uk) or contact:
External Communications Manager
Tyndall Centre for Climate Change Research
University of East Anglia, Norwich NR4 7TJ, UK
Phone: +44 (0) 1603 59 3906; Fax: +44 (0) 1603 59 3901
Email: tyndall@uea.ac.uk
Other titles in the Tyndall Working Paper series include:

1. A country-by-country analysis of past and future warming rates, November 2000


2. Integrated Assessment Models, March 2001
3. Socio-economic futures in climate change impact assessment: using scenarios as
‘learning machines’, July 2001
4. How high are the costs of Kyoto for the US economy?, July 2001
5. The issue of ‘Adverse Effects and the Impacts of Response Measures’ in the
UNFCCC, July 2001
6. The identification and evaluation of suitable scenario development methods for the
estimation of future probabilities of extreme weather events, July 2001
7. Security and Climate Change, October 2001
8. Social Capital and Climate Change, October 2001
9. Climate Dangers and Atoll Countries, October 2001
10. Burying Carbon under the Sea: An initial Exploration of Public Opinions, December
2001
11. Representing the Integrated Assessment of Climate Change, Adaptation and
Mitigation, December 2001
12. The climate regime from The Hague to Marrakech: Saving or sinking the Kyoto
Protocol?, December 2001
13. Technological Change, Industry Structure and the Environment, January 2002
14. The Use of Integrated Assessment: An Institutional Analysis Perspective, April 2002
15. Long run technical change in an energy-environment-economy (E3) model for an IA
system: A model of Kondratiev waves, April 2002
16. Adaptation to climate change: Setting the Agenda for Development Policy and
Research, April 2002
17. Hydrogen energy technology, April 2002
18. The development of large technical systems: implications for hydrogen, March 2002
19. The role of hydrogen in powering road transport, April 2002
20. Reviewing organisational use of scenarios: case study - evaluating UK energy policy
options, August 2002

The Tyndall Working Papers are available online at:


http://www.tyndall.ac.uk/publications/working_papers/working_papers.shtml

S-ar putea să vă placă și