30.01.2015 Views

Version Control - Granta Design

Version Control - Granta Design

Version Control - Granta Design

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Materials for Energy<br />

Storage Systems<br />

Materials for Energy Storage<br />

Systems—A White Paper<br />

Mike F Ashby and James Polyblank<br />

Engineering Department, University of Cambridge, Trumpington Street, Cambridge, CB2 1PZ, UK<br />

<strong>Version</strong> 2.0, first published January 2012<br />

© <strong>Granta</strong> <strong>Design</strong>, 2012<br />

Image of flywheel for regenerative braking, courtesy of BP Research Centre, Sunbury, UK.<br />

Materials for Energy Storage Systems © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Table of Contents<br />

1. Introduction—the need for energy storage systems ................................................................1<br />

2. Overview—selecting energy storage systems .........................................................................2<br />

3. Pumped hydro storage ...........................................................................................................11<br />

4. Compressed air energy storage (CAES) ...............................................................................12<br />

5. Springs ...................................................................................................................................14<br />

6. Flywheels ...............................................................................................................................16<br />

7. Thermal Storage ....................................................................................................................17<br />

8. Batteries .................................................................................................................................18<br />

9. Hydrogen Energy Storage .....................................................................................................23<br />

10. Capacitors and super-capacitors .........................................................................................24<br />

11. Superconducting Magnetic Energy Storage (SMES)...........................................................26<br />

Appendix 1 Definition of terms ...................................................................................................27<br />

Appendix 2: Approximate Material Intensities for Energy Storage Systems ..............................29<br />

Further Reading .........................................................................................................................34<br />

Materials for Energy Storage Systems © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


1. Introduction—the need for<br />

energy storage systems<br />

Energy storage systems are the source of<br />

power for products as diverse as hearing<br />

aids and satellites. Two sectors have a<br />

particular interest in large-scale energy<br />

storage systems: the utilities industry and<br />

the automotive industry. The portable<br />

electronics industry and the military also<br />

require energy storage systems, but on a<br />

smaller scale.<br />

The utility industry is faced with the<br />

problem that energy demand is not<br />

constant but fluctuates through the day.<br />

With the increasing dependence on<br />

renewable energy sources (wind, wave,<br />

solar), energy supply, too, fluctuates.<br />

Energy storage systems ensure that,<br />

when supply is high and demand is low,<br />

energy is not wasted, and that when<br />

supply is low and demand is high, the<br />

demand can still be met. An electricity grid<br />

allows electricity to be generated at one<br />

location and used at another. Energy<br />

storage systems allow electricity to be<br />

generated at one time and used at<br />

another. The automotive industry has<br />

rather different needs. Legislation to<br />

reduce emissions is driving the<br />

development of a new generation of low<br />

carbon vehicles. Gasoline is an<br />

exceptionally effective form of portable<br />

energy. Replacing it requires a storage<br />

system that can provide acceptable range<br />

and allow acceptably fast refueling or<br />

recharging.<br />

This report surveys prominent energy<br />

storage systems (Table 1). Section 2<br />

provides an overview of the findings and<br />

describes how energy storage systems<br />

can be selected. Subsequent sections<br />

provide a deeper analysis of each system,<br />

with a particular emphasis on their<br />

ultimate limits and the demands they place<br />

on material supply.<br />

Table 1. Classes and members of energy storage systems.<br />

Class<br />

Hydrocarbons (for comparison)<br />

Mechanical storage<br />

Thermal storage<br />

Chemical storage<br />

Electro-magnetic storage<br />

Systems<br />

Conventional fuels<br />

Pumped hydro<br />

Compressed air energy storage<br />

Springs<br />

Flywheels<br />

Thermal storage<br />

Li-ion batteries<br />

Sodium-sulfur batteries<br />

Lead-acid batteries<br />

Nickel cadmium batteries<br />

Nickel-metal hydride batteries<br />

Vanadium flow batteries<br />

Hydrogen fuel cells<br />

Super capacitors<br />

Superconducting magnetic energy storage<br />

Materials for Energy Storage Systems 1 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


As we found with systems for low carbon<br />

power generation (see Materials for Low-<br />

Carbon Power—A White Paper), the<br />

performance of energy storage systems<br />

depends on the type and scale of the<br />

system, its location, and the way it is<br />

managed. Technical developments<br />

continue to improve the performance of<br />

existing systems and add new ones. The<br />

figures and tables of this paper show<br />

representative ranges, but there is no<br />

guarantee that these ranges enclose all<br />

members of a given system.<br />

2. Overview—selecting energy<br />

storage systems<br />

Even within the utility industry and the<br />

automotive industry, there are different<br />

applications of energy storage devices.<br />

The utility industry needs short bursts of<br />

energy at high power for frequency<br />

regulation (that is, maintaining a constant<br />

mains frequency), while they need more<br />

sustained power for energy management<br />

(ensuring that demand is met, and supply<br />

is not wasted). Similarly, the automotive<br />

industry needs short bursts of energy for<br />

acceleration and more sustained power for<br />

long distance cruising.<br />

Six performance metrics for energy<br />

storage systems appear in Table 2.<br />

Specific energy and energy density are<br />

the energy storage capacity per unit mass<br />

and unit volume respectively. The<br />

operating cost is the approximate cost of<br />

one charge/discharge cycle of one unit of<br />

energy, and includes the cost of<br />

maintenance, heating, labor, etc. Specific<br />

power is the discharge power capacity per<br />

unit mass. The efficiency is defined as the<br />

ratio of total energy outputted by the<br />

system over total energy put into the<br />

system in one charge/discharge cycle.<br />

The cycle life is defined as the number of<br />

charge/discharge cycles possible until the<br />

capacity of the storage system drops to<br />

80% of its initial capacity.<br />

Figure 1 identifies the storage systems<br />

which are best for short power bursts and<br />

those able to provide more sustained<br />

power. The axes are Specific power<br />

(power per unit mass) and Specific energy<br />

(energy storage capacity per unit mass),<br />

with contours of constant discharge time.<br />

Superconducting magnetic energy storage<br />

(SMES), electric double layer capacitors<br />

Figure 1. A plot of power density against energy density of energy storage systems,<br />

with lines of charge/discharge time.<br />

Materials for Energy Storage Systems 2 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


(EDLC, also known as super-capacitors or<br />

ultra-capacitors), and flywheels can<br />

provide short bursts of power, while<br />

pumped hydro storage and thermal<br />

storage are better suited for longer<br />

duration power supply.<br />

Table 2. Performance metrics of energy storage systems 1 .<br />

Once a selection of storage systems with<br />

suitable discharge times have been<br />

identified for a given application, it then<br />

remains to select the energy storage<br />

system which has the best performance,<br />

and the lowest resource intensities.<br />

Storage System<br />

Specific<br />

Energy<br />

Energy<br />

Density<br />

Operating<br />

Cost<br />

Specific<br />

Power<br />

Efficiency<br />

Cycle Life<br />

(MJ/kg) (MJ/m 3 ) ($/MJ) (W/kg) (%) (#-cycles)<br />

Conventional fuels 20-50 15,000-<br />

72,000<br />

- - 18-50 1<br />

Pumped hydro 0.002-0.005 2-5 0.0006-<br />

0.0014<br />

0.02-0.3 70-80 400,000<br />

Compressed air energy<br />

storage<br />

0.36 25 0.0001-<br />

0.0019<br />

8 65-70 15,000<br />

Springs 0.00014-<br />

0.00033<br />

0.6-1.1 - - 98-99.9 Depends<br />

on loading 2<br />

Flywheels 0.002-0.025 2.1 0.0008-<br />

0.0017<br />

100-<br />

10,000<br />

75-85 150,000<br />

Thermal storage 0.032-0.036 160 0.0008-<br />

0.0019<br />

Li-ion batteries 0.32-0.68 720-1,400 0.0019-<br />

0.0047<br />

1.5-1.7 72-85 10,000-<br />

15,000<br />

250-340 80-90 300-2,000<br />

Sodium-sulfur batteries 0.2-0.7 140-540 0.0039 10-15 75-83 3600-4700<br />

Lead-acid batteries 0.07-0.18 200-430 0.0008-<br />

0.0028<br />

4-180 70-90 200-1500<br />

Nickel cadmium<br />

batteries<br />

Nickel-metal hydride<br />

batteries<br />

0.08-0.23 72-310 0.0008-<br />

0.0055<br />

0.1-0.43 190-1300 0.0006-<br />

0.0028<br />

30-150 60-85 800-1200<br />

4-140 65-85 300-1000<br />

Vanadium flow batteries 0.07-0.13 110-170 0.0039 2.2-3.1 71-88 10000-<br />

16000<br />

Hydrogen fuel cells 0.02-0.8 1-70 0.0006-<br />

0.0041<br />

5-20 27-35 5,000-<br />

10,000<br />

Super capacitors 0.010-0.020 10-25 0.0008-<br />

0.0019<br />

1,000-<br />

20,000<br />

90-95 1,000,000<br />

Superconducting<br />

magnetic energy<br />

storage<br />

0.001-0.02 2-10 0.0003-<br />

0.0083<br />

500-<br />

20,000<br />

85-92 50,000-<br />

200,000<br />

1 Some of the data in this table are easy to find, but others are not. Some have to be estimated from diagrams or<br />

schematics of the system, some deduced by analogy with similar systems, and some inferred from the physics<br />

on which the system depends. The values vary greatly with design, location and scale of the system, allowing<br />

wide variation. That means precision is low, but the difference between the values of competing systems is<br />

sufficiently great that it is still possible to draw meaningful conclusions.<br />

2 The cycle life of a spring depends on loading—for low alloy steels, for example, an increase in loading of 50%<br />

can reduce the lifetime by six orders of magnitude.<br />

Materials for Energy Storage Systems 3 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Figure 2. Specific energy and energy density of energy storage methods, including a number of fuels for<br />

comparison. The densities of most solids sit within an order of magnitude of that of water (1,000kg/m 3 ), so it is<br />

no surprise that physical density of all the storage systems (apart from gases) sit within this range.<br />

The importance of each performance<br />

metric depends on the application. The<br />

automotive industry requires energy<br />

storage systems that are small and light,<br />

so are interested in high Specific energy<br />

and Energy density. These two metrics<br />

are the axes of Figure 2. Compared to<br />

fuels, with specific energies from 20 MJ/kg<br />

(for methanol) to 50 MJ/kg (for LPG) and<br />

energy densities from 16,000 MJ/m 3 (for<br />

methanol) and 72,000 MJ/m 3 (for coal), all<br />

of the “low carbon” energy storage<br />

systems perform poorly. Their specific<br />

energies and energy densities are at least<br />

one order of magnitude smaller than those<br />

of hydrocarbon fuels. Even explosives<br />

(which must include both reactants,<br />

whereas the masses of fuels do not<br />

include the oxygen the fuel reacts with)<br />

store much more energy than any other<br />

storage system per unit mass or volume.<br />

Many of these systems are not yet fully<br />

mature; future research and development<br />

will certainly improve their performance.<br />

But there are upper limits, physical and<br />

practical, beyond which performance<br />

cannot be pushed. A plot of these upper<br />

limits for specific energy and energy<br />

density (Figure 3) shows how they might<br />

compete with fuels in the future. The<br />

assumptions used to estimate these<br />

values are discussed in later sections.<br />

They are extremely optimistic. The ratio of<br />

the current specific energy to the<br />

estimated upper limit of each system<br />

(Figure 4) is around 1% for all except<br />

CAES and SMES, suggesting that there is<br />

some consistency in the level of optimism<br />

that has been applied. With the most<br />

optimistic assumptions, the only energy<br />

storage systems that approach the energy<br />

densities of fuels are hypothetical<br />

flywheels made from carbon nano-tubes<br />

(CNTs), hypothetical lithium-fluorine<br />

batteries, and superconducting magnetic<br />

energy storage (SMES). Even<br />

approaching these limits is improbable<br />

because of practicalities discussed in the<br />

subsequent sections of the report. It is<br />

very hard to compete successfully with<br />

hydrocarbon fuels.<br />

Materials for Energy Storage Systems 4 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Figure 3. A plot of energy mass density vs. energy volume density optimistically<br />

achievable by various technologies.<br />

Figure 4. The ratio of the current specific energy to the limiting specific energy of<br />

energy storage systems.<br />

Materials for Energy Storage Systems 5 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Figure 5 and Figure 6 show bar-charts of<br />

cycle efficiency and cycle life. Poor<br />

efficiency means large losses, making<br />

storage expensive. Poor cycle life makes<br />

investment in the system less attractive<br />

because of the cost and inconvenience of<br />

replacement.<br />

Figure 5. Cycle efficiency of energy storage systems.<br />

Fossil fuel engines/power plants are included for comparison.<br />

Figure 6. Cycle life of energy storage systems.<br />

Materials for Energy Storage Systems 6 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Constructing an energy storage system<br />

requires resources. Five resource<br />

intensities of construction appear in<br />

Table 3. All are quoted per unit of energystorage<br />

capacity. Capital intensity is the<br />

cost of construction per MJ of storage<br />

capacity. Area intensity is the typical area<br />

used to site the storage system, per MJ of<br />

storage capacity. Material intensity is the<br />

mass of raw materials required to<br />

construct the storage system, per MJ.<br />

Energy intensity is the primary energy<br />

required to construct the storage system,<br />

and CO 2 intensity is the CO 2 emitted in<br />

constructing the storage system, per MJ of<br />

storage capacity.<br />

.<br />

Table 3. Resource intensities of construction of energy storage devices 3<br />

Storage System<br />

Capital<br />

Intensity<br />

Area Intensity<br />

Material<br />

Intensity<br />

Energy<br />

Intensity<br />

CO 2 Intensity<br />

($/MJ) (m 2 /MJ) (kg/MJ) (MJ embodied /MJ) (kg-CO 2 /MJ)<br />

Conventional<br />

Fuels<br />

0.045-0.065 4 - - 0.1-0.2 4 0.2-0.4 4<br />

Pumped Hydro 45-120 0.02-0.08 60-120 100-200 8-16<br />

Compressed Air<br />

Energy Storage<br />

4-20 0.008 2-12 74 5.3<br />

Springs 500-1,600 - 3,000-7,100 340,000-<br />

550,000<br />

24,000-42,000<br />

Flywheels 400-7,000 0.08-0.2 17-500 750-760 90-100<br />

Thermal Storage 14-26 0.003 0.4-50 120-130 8.9-9.0<br />

Li-ion Batteries 140-440 0.002-0.008 1.5-2.7 330-580 19-50<br />

Sodium-Sulphur<br />

Batteries<br />

Lead-Acid<br />

Batteries<br />

Nickel Cadmium<br />

Batteries<br />

Nickel-Metal<br />

Hydride Batteries<br />

Vanadium Flow<br />

Batteries<br />

Hydrogen Fuel<br />

Cells<br />

28-280 0.005-0.008 1.4-5 360-640 30-50<br />

50-220 0.009-0.03 4.5-12 110-980 5-130<br />

200-330 0.006-0.03 3.5-10 390-640 28-47<br />

130-440 0.003-0.03 2-7 550-940 28-67<br />

100-300 0.01 3.8-7 170-180 25-26<br />

55-2,300 0.001-0.015 1.3-50 140-150 9.7-9.8<br />

Super-capacitors 14,700-41,000 0.11-0.13 45-65 3,700-6,500 210-360<br />

Superconducting<br />

Magnetic Storage<br />

30,000-260,000 0.25-7.1 50-1,000 1,600-1,700 240-250<br />

3 Some of the data in this table are easy to find, but others are not. Some have to be estimated from diagrams or<br />

schematics of the system, some deduced by analogy with similar systems, and some inferred from the physics<br />

on which the system depends. The values vary greatly with design, location, and scale of the system, allowing<br />

wide variation. That means precision is low, but the difference between the values of competing systems is<br />

sufficiently great that it is still possible to draw meaningful conclusions.<br />

4 Note that conventional fuels can only be used once, so these figures cannot be compared directly with the other<br />

systems, but are included for completeness.<br />

Materials for Energy Storage Systems 7 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Figure 7. The capital and energy intensities of construction of energy storage systems,<br />

both per cycle over the lifetime of the system.<br />

The capital and energy required to install a<br />

given energy-storage capacity depends on<br />

capital and energy intensities of the<br />

storage system. Figure 7 illustrates this. It<br />

has axes of<br />

Construction energy per MJ of<br />

Cycle life<br />

and<br />

Construction capital per MJ of<br />

Cycle life<br />

storage capactiy<br />

storage capactiy<br />

Compressed air energy storage has the<br />

lowest capital intensity per MJ-cycle.<br />

Super-capacitors are the least energy<br />

intensive storage system, largely because<br />

of their high cycle life. The energy and<br />

capital intensities of batteries are inflated<br />

by their lower cycle life, making some of<br />

them more expensive and more energy<br />

intensive per cycle than gasoline.<br />

Investment decisions for energy storage<br />

are influenced by the total cost per unit<br />

storage capacity per cycle. This is:<br />

Average cost ($ / MJ − cycle )<br />

= Construction cost ($ / MJ )<br />

+ Operating cost ($/<br />

MJ − cycle )<br />

Cycle life<br />

Figure 8 plots this for each system.<br />

It reveals that CAES is still the cheapest.<br />

Large scale implementations of energy<br />

storage systems for grid stability makes<br />

considerable demands on material supply<br />

and land area, making Material intensity<br />

and Area intensity metrics of particular<br />

importance to the utility industry. They are<br />

plotted in Figure 9. Thermal storage,<br />

lithium-ion batteries, and hydrogen<br />

storage use the least area and materials.<br />

Materials for Energy Storage Systems 8 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Figure 8. Total costs of energy storage systems per unit storage capacity per cycle.<br />

Figure 9. The area and material intensities of energy storage systems.<br />

Using an overall material intensity is, of<br />

course, an oversimplification. What is<br />

important is which materials are used, and<br />

whether they are “critical”. The bills of<br />

materials for the energy storage systems<br />

studied here are listed in Appendix 2.<br />

These allow a comparison between the<br />

demands of each system and the current<br />

world production of critical materials,<br />

highlighting where supply might be a<br />

problem.<br />

Materials for Energy Storage Systems 9 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


What is a critical material There are four<br />

reasons for a material to be categorized in<br />

this way:<br />

• Production is primarily from<br />

sources considered unreliable for<br />

geo-political reasons;<br />

• Lack of substitutes for the material<br />

in its main application;<br />

• Supply is limited due to the<br />

economics of its production; or<br />

• Extreme price volatility.<br />

The annual productions of 27 materials<br />

that meet one or more of these criteria are<br />

shown in Figure 10, color-coded to show<br />

the reasons for concern.<br />

In subsequent sections we examine each<br />

energy storage system in turn, paying<br />

particular attention to critical materials. To<br />

do this we consider a hypothetical<br />

scenario, one that is compatible with that<br />

used in Materials for Low-Carbon Power—<br />

A White Paper. To complement the<br />

2,000 GW of power generation capacity<br />

hypothetically being replaced over a 10<br />

year period, 10 hours of storage capacity<br />

will be installed. This will require a storage<br />

capacity of 20,000 GWh, or 72 PJ<br />

(petajoule = 10 15 J). From this we calculate<br />

the fraction of current world production of<br />

each critical material that would be<br />

required to build the installations,<br />

revealing where material supply might be<br />

a problem.<br />

Figure 10. Current (2010) annual world production of 27 critical materials, highlighting the<br />

reasons for their being identified as critical.<br />

Materials for Energy Storage Systems 10 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


3. Pumped hydro storage<br />

Energy is stored if water is pumped from a<br />

lower reservoir to an upper reservoir that<br />

is 100-1,000 m higher (Figure 11). Energy<br />

is recovered by allowing the water to<br />

return to the lower reservoir through a<br />

generator like that of a hydro-power plant.<br />

The gravitational potential energy stored in<br />

a mass m of water by pumping it to a<br />

height h is:<br />

W = m g h<br />

(1)<br />

where g is the acceleration due to gravity<br />

(10 m/s 2 ). Assuming that the mass and<br />

volume of the dam and ancillary<br />

equipment are negligible compared to<br />

those of the water and that energy is<br />

recovered with 100% efficiency, the<br />

specific energy and the energy density are<br />

W W m<br />

= g h and = g h = ρ g h (2)<br />

m<br />

V V<br />

where V is the volume of water and ρ is<br />

the density of water (1,000 kg/m 3 ). The<br />

quantities ρ and g are approximately<br />

constant on earth, so the specific energy<br />

and energy density of a pumped hydro<br />

system depend only on h —the height of<br />

the upper reservoir above the lower one.<br />

To put an absolute upper limit on this, we<br />

postulate a pumped hydro system with the<br />

upper reservoir at the highest point on<br />

earth (Everest at 8,850m), and the lower<br />

reservoir at sea level, giving the<br />

(unattainable) limits<br />

and<br />

*<br />

⎛W<br />

⎞<br />

⎜ ⎟<br />

⎝ m ⎠<br />

*<br />

⎛ W ⎞<br />

⎜ ⎟<br />

⎝ V ⎠<br />

= 10 x 8850 J / kg<br />

= 1000 x 10 x 8850 = 88 , 500<br />

= 0.09 MJ / kg<br />

, 000<br />

3<br />

= 90 MJ / m<br />

J / m<br />

3<br />

Figure 11. Diagram of a pumped hydro storage system.<br />

Materials for Energy Storage Systems 11 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Even this fantasy scenario gives values of<br />

specific energy and energy density that<br />

are almost three orders of magnitude<br />

smaller than those of oil. Real systems<br />

have values that are between 10 and 100<br />

times smaller still.<br />

The simplicity of pumped hydro systems<br />

makes them the preferred method of<br />

utility-scale energy storage, accounting for<br />

99% of world-wide capacity. An<br />

approximate bill of materials for such a<br />

system appears in Appendix 2. It suggests<br />

that the material demands for large scale<br />

deployment (following the scenario<br />

described in Section 2) would be<br />

manageable, requiring only 0.2% of the<br />

world production of chromium and<br />

manganese, the only critical materials<br />

required. The greatest obstacle to<br />

expansion is simply that of finding useable<br />

locations.<br />

4. Compressed air energy<br />

storage (CAES)<br />

CAES systems store energy by<br />

compressing air. Energy is recovered by<br />

expanding the air through a pneumatic<br />

motor or turbine connected to a generator.<br />

The compressed air can be stored in small<br />

tanks for mobile (e.g., automotive)<br />

applications, or in underground caverns or<br />

large balloons under the sea (which keeps<br />

it at high pressure) for utility scale<br />

applications.<br />

When air is compressed its temperature<br />

rises. If this heat leaks away, energy is<br />

lost. When the air is expanded again, it<br />

cools and can freeze the turbine. There<br />

are two ways of dealing with this. One is to<br />

insulate the compressed air chamber so<br />

that the heat does not escape, allowing<br />

adiabatic compression. The other is to<br />

pump slowly and allow the heat to escape<br />

or to be stored separately so that the<br />

temperature of the compressed air<br />

chamber remains constant, giving<br />

isothermal compression. Isothermal<br />

Figure 12. Diagram of a hybrid CAES system.<br />

Materials for Energy Storage Systems 12 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


compression must be followed by slow,<br />

isothermal expansion, allowing the heat to<br />

return to the air. In theory, either method<br />

could be 100% efficient, but in practice an<br />

efficiency of about 70% is the best<br />

achievable.<br />

Today’s CAES systems combine<br />

compressed air with natural gas. The gas<br />

burns to heat the air before the mixture<br />

enters the turbine (Figure 12). In a<br />

conventional gas power plant, substantial<br />

energy is used to compress air. The<br />

stored compressed air removes the need<br />

for this, increasing the efficiency. Taking<br />

out the contribution of the gas, CAES<br />

systems have an efficiency of 70% for<br />

round-trip compressed air storage.<br />

Large scale deployment of CAES systems<br />

would be manageable, requiring only<br />

using 0.001% of the annual production of<br />

chromium, and 0.3% of manganese<br />

(Appendix 2). Mobile applications require<br />

a pressure vessel, but this is unlikely to<br />

put pressure on supply of critical<br />

materials.<br />

The specific energy of today’s CAES<br />

systems is about 0.36 MJ/kg, and the<br />

energy density is about 25 MJ/m 3 , barely<br />

1% of the values for hydrocarbon fuels.<br />

Could they be improved The energy<br />

stored by isentropically compressing<br />

n moles of air at temperature T from<br />

pressure p A to p B into a pressure vessel of<br />

volume V is:<br />

WA−B<br />

⎛ p ⎞<br />

⎜ B<br />

= nR T ln<br />

⎟ =<br />

⎝ p A ⎠<br />

⎛ p ⎞<br />

⎜ B<br />

pB<br />

V ln<br />

⎟<br />

⎝ p A ⎠<br />

(3)<br />

The mass of a thin walled spherical vessel<br />

of radius, r, thickness, t, and made out of<br />

material of density, ρ, is:<br />

m = 4π<br />

r<br />

2<br />

t ρ<br />

(4)<br />

The wall thickness required to ensure that<br />

the wall does not yield under a pressure<br />

difference ∆p is:<br />

where<br />

1 ∆p r<br />

t ≥ 2 σ<br />

σ y<br />

(5)<br />

is the yield strength of the wall<br />

material. Substituting this into equation (5)<br />

gives:<br />

m =<br />

(6)<br />

The volume of the pressure vessel is<br />

4 3<br />

V = π r , allowing this equation to be reexpressed<br />

3<br />

as:<br />

m =<br />

(7)<br />

To minimize the mass, we need to<br />

maximize σ y / ρ . The structural material<br />

with the largest value of this index is<br />

CRFP with / ρ = 1.6 MJ/kg. Noting that,<br />

at high pressures<br />

the vessel is:<br />

y<br />

3<br />

2π<br />

r ρ ∆p<br />

σ<br />

y<br />

3 V ∆p<br />

2<br />

σ y<br />

ρ<br />

σ<br />

y<br />

∆p ≈<br />

mvessel = 0. 94 V p B<br />

pB<br />

the mass of<br />

(8)<br />

where pressures are in units of MPa and<br />

volume in m 3 . The mass of the air itself<br />

(at 300K):<br />

pB<br />

mair = ρ air V = 12 pBV<br />

(9)<br />

p A<br />

where ρ air is the density of air at<br />

atmospheric pressure (1.2 kg/m 3 at<br />

0.1 MPa). The specific energy is thus:<br />

m<br />

vessel<br />

W<br />

+ m<br />

air<br />

=<br />

p<br />

0 . 94 p<br />

p B<br />

V ln( )<br />

p A<br />

V + 12 p V<br />

p B<br />

= 0 . 08 ln( ) MJ/kg<br />

p A<br />

(10)<br />

It only remains to select the compression<br />

pressure. The most powerful compressors<br />

available today typically compress from<br />

atmospheric pressure (0.1 MPa) to<br />

100 MPa. (These compressors are used<br />

for hot isostatic pressing for powder<br />

B<br />

B<br />

B<br />

Materials for Energy Storage Systems 13 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


processing of materials.) This gives a<br />

specific energy of 0.55 MJ/kg, and an<br />

energy density of 580 MJ/m 3 . Even<br />

without the pressure vessel (e.g., in a<br />

solution mined salt cavern), the specific<br />

energy would be 0.59 MJ/kg. This is still<br />

two orders of magnitude smaller than<br />

conventional fuels.<br />

It should also be noted that, in calculating<br />

this upper limit, we have applied no safety<br />

factor (it should be at least 3), neglected<br />

the low efficiency of pneumatic pumps<br />

(20 – 30%), and assumed 100% energy<br />

recovery (in reality 80% at best). It is clear<br />

that compressed air storage can work for<br />

utility power but for mobile applications it<br />

cannot compete with fossil fuels.<br />

5. Springs<br />

Springs store energy by compressing or<br />

extending material elastically. There are<br />

no examples of springs being used for<br />

utility scale energy storage or to provide<br />

propulsion in automobiles. This is because<br />

they have very low specific energy. We<br />

highlight this here by considering the<br />

maximum energy we could store in a<br />

spring using today’s materials.<br />

The specific energy, energy density, and<br />

the capital intensity of energy storage in a<br />

material when stretched uniaxially to yield<br />

are:<br />

Energy density<br />

Specific energy<br />

Capital intensity<br />

W 1 σ y 2<br />

=<br />

V 2 E<br />

W<br />

m<br />

σ y<br />

2<br />

ρ E<br />

(11,a)<br />

(11,b)<br />

(11,c)<br />

where W is energy stored, V is the<br />

volume of material, m is the mass of<br />

material, C is the total cost of material,<br />

is the yield stress of the material, E is<br />

σ y<br />

the Young’s modulus, ρ is the density,<br />

and C m is the material cost per unit mass.<br />

Table 4 lists values for five different<br />

materials.<br />

C<br />

W<br />

=<br />

1<br />

2<br />

E ρ C<br />

= 2 m<br />

2<br />

σ y<br />

Figure 13. A titanium spring.<br />

Image courtesy of G and O Springs Ltd.<br />

Table 4. The performance of various materials for springs.<br />

Material MJ/m 3 MJ/kg USD/MJ Cycle<br />

Efficiency<br />

Polyurethane rubber (Unfilled) 30-400 0.03-0.4 20-200 40-70%<br />

PEBA (Shore D25) 21-38 0.021-0.038 220-400 70-94%<br />

Epoxy/S-glass fiber (0° ply) 30-33 0.015-0.017 1500-1700 98%<br />

Low alloy steel, AISI 9255,<br />

tempered 205°C<br />

8-12 0.001-0.0015 550-830 99.8-99.9%<br />

Carbon nano-tubes 8,400 5 30,000-160,000<br />

Materials for Energy Storage Systems 14 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Polyurethane rubber is the best according<br />

to all three of these figures of merit.<br />

However cyclically compressing and<br />

expanding polyurethane rubber is<br />

extremely lossy, with a mechanical loss<br />

coefficient of tan(δ)=0.05-0.1, leading to a<br />

cycle efficiency of 40-70%. PEBA (Shore<br />

D25) has a better material efficiency of<br />

94%, but has a lower specific energy of<br />

0.021-0.038 MJ/kg compared to 0.03-<br />

0.4 MJ/kg for polyurethane rubber.<br />

Epoxy/S-glass fiber makes the lightest and<br />

smallest spring, while low alloy steel<br />

makes the cheapest. These have much<br />

higher cycle efficiencies (98% and 99.9%<br />

respectively), but the overall efficiency will<br />

be reduced by the mechanisms used to<br />

store and extract the energy. Thus the<br />

energy density of springs made from<br />

today’s materials is three orders of<br />

magnitude smaller than conventional<br />

fuels.<br />

Hill et al. (2009) claim that it would be<br />

possible to store 8,400 MJ/m 3 or 5 MJ/kg<br />

in carbon nano-tube springs. This might<br />

be true of a single nanotube, but scaling<br />

up to any useful size requires that the<br />

nanotubes are bonded in some kind of<br />

matrix, reducing the specific energy. A<br />

more realistic upper bound is to postulate<br />

that the spring material could be loaded to<br />

the theoretical strength, roughly E / 10 .<br />

Then using (as an example) the modulus<br />

and density of steel (200 GPa and<br />

7,900 kg/m 3 ) we find upper bounds from<br />

equations 11,a and 11,b as follows.<br />

Energy density:<br />

W<br />

V<br />

=<br />

2<br />

1 σ y<br />

2 E<br />

Specific energy:<br />

W<br />

m<br />

=<br />

2<br />

=<br />

1 σ y<br />

2 ρ E<br />

=<br />

1<br />

2<br />

E<br />

100<br />

E<br />

100<br />

3<br />

= 1000 MJ / m<br />

These are still far below the values for<br />

hydrocarbons.<br />

1<br />

2<br />

0.13 MJ / kg.<br />

ρ<br />

Figure 14. Diagram of a flywheel. The motor/generator is adapted to fit inside the flywheel to save space.<br />

The system is buried underground for safety.<br />

Materials for Energy Storage Systems 15 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


6. Flywheels<br />

A spinning flywheel stores kinetic energy<br />

(Figure 14), which is recovered by<br />

coupling the flywheel to a generator. An<br />

efficient flywheel is designed to have as<br />

large a rotational moment of inertia as<br />

possible and to spin fast. Frictional losses<br />

in the bearings and surrounding air reduce<br />

efficiency. They are reduced by using<br />

superconducting magnetic bearings to<br />

levitate the rotor and by cooling or<br />

pumping the air out of the containment. All<br />

of these require energy, so they too<br />

reduce efficiency slightly.<br />

Present-day flywheels have a specific<br />

energy in the range 0.002 – 0.025 MJ/kg<br />

and an energy density is between 1.7 and<br />

23 MJ/m 3 when ancillary devices are<br />

included, values that are much smaller<br />

than hydrocarbon fuels. What are the<br />

upper limits for both The following<br />

estimate gives an idea.<br />

The kinetic energy of a rotating cylinder is:<br />

1 W = I ω<br />

2<br />

(12)<br />

2<br />

where I is the rotational moment of<br />

inertia and ω the angular velocity. The<br />

moment of inertia is maximized for a given<br />

mass of material by making it into a thinwalled<br />

tube. If the tube has mass m ,<br />

radius R and wall thickness t<br />

(Figure 15, left), its moment of inertia per<br />

unit length is:<br />

2<br />

I = m R<br />

(13)<br />

Equating the centrifugal load in a small<br />

element of this ring to the resolved<br />

component of the circumferential stress in<br />

the wall when this stress is just below the<br />

its yield strength gives the maximum<br />

allowable angular velocity, ω :<br />

2<br />

from which<br />

σ y<br />

dmω R = ( R t ρ dθ<br />

) ω R = t dθ<br />

σ<br />

ω<br />

2<br />

giving a specific energy of<br />

W<br />

m<br />

σ y<br />

=<br />

2<br />

R ρ<br />

=<br />

y<br />

1 σ y<br />

2 ρ<br />

(14)<br />

MJ/kg (15)<br />

where σ is in MPa and the density ρ is<br />

in kg/m 3 . Among today’s materials, CFRP<br />

has the highest value of σ y / ρ ≈<br />

1.6 MJ/kg, allowing a flywheel with specific<br />

energy of 0.8 MJ/kg. This is two orders of<br />

magnitude smaller than the value for<br />

hydrocarbons, and we have not included<br />

the protective burst shield, the motorgenerator,<br />

or any safety factors (up to five<br />

for composite materials)—all of which<br />

reduce the specific energy.<br />

2<br />

y<br />

Figure 15. Left: A flywheel. The maximum kinetic energy it can store is limited by its strength.<br />

Right: Demand ratios for critical materials used in flywheel energy storage systems with superconducting<br />

magnetic bearings. *CFRP is included because, even though it is not a critical material, large scale<br />

implementation is likely to stretch its supply.<br />

Materials for Energy Storage Systems 16 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


It has been suggested that flywheels might<br />

be constructed that exploited the great<br />

strength ( = 100 GPa, Peng et al.,<br />

σ y<br />

2008) and low density ( ρ = 1,300 kg/m 3 ,<br />

Collins & Avouris, 2000) of carbon nanotubes,<br />

allowing a specific energy of<br />

40 MJ/kg, comparable with conventional<br />

fuels. However, this neglects the need for<br />

a supporting matrix. A more realistic upper<br />

limit, following the reasoning of the<br />

previous section, is to imagine a material<br />

with the density and modulus of steel and<br />

with the theoretical strength E / 10. , giving<br />

a specific energy<br />

W<br />

m<br />

=<br />

1 σ y<br />

2 ρ<br />

=<br />

1<br />

2<br />

E<br />

10<br />

=<br />

ρ<br />

1.3<br />

MJ/kg<br />

Large scale implementation of flywheel<br />

storage systems with superconducting<br />

bearings using yttrium barium copper<br />

oxide superconductors and CFRP disks<br />

(Appendix 2) could exert pressure on<br />

supply of yttrium and CFRP (Figure 15,<br />

right).<br />

7. Thermal Storage<br />

Thermal energy storage takes two forms.<br />

1. Heating an inert solid or liquid,<br />

using its heat capacity (specific<br />

heat), (J/kg.K) to retain energy,<br />

C p<br />

2. Causing a solid to melt, using its<br />

latent heat of fusion, L m (J/kg) to<br />

capture energy.<br />

Energy is recovered by passing a heattransfer<br />

fluid through the hot or molten<br />

body, passing the heat to a heat<br />

exchanger where it generates steam or<br />

hot gas for space heating or to drive a<br />

turbine. Utility-scale thermal energy<br />

storage is used to smooth the power<br />

output of concentrated solar plants<br />

(Figure 16). The sun’s radiation is focused<br />

on a collector containing molten salt<br />

(typically 40% potassium nitrate, 60%<br />

sodium nitrate). The molten salt is held in<br />

hot tanks until the energy is needed when<br />

an oil-loop from the hot tank carries the<br />

heat to a steam generator and turbine to<br />

generate electrical power.<br />

Figure 16. Diagram of a molten salt thermal storage system connected to a solar tower.<br />

Only the thermal storage block is included when considering material usage.<br />

Materials for Energy Storage Systems 17 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Table 5. Properties of materials for thermal energy storage.<br />

Material<br />

Melting point<br />

( o C)<br />

Specific heat<br />

(kJ/kg o C)<br />

Energy density<br />

(MJ/m 3 o C)<br />

Latent heat of<br />

fusion (MJ/kg)<br />

Aluminum 660 0.92 2.5 0.39<br />

Cast Iron 1,200 0.54 3.9 0.27<br />

Lead 324 0.13 1.5 0.034<br />

Lithium 180 3.6 1.9 0.43<br />

Sodium 100-760 1.3 0.95 0.11<br />

Molten alt (50% KNO 3 ; 40% NaNO 2 ) 140-540 1.6 2.6 -<br />

Fireclay 1450 1.0 2.1-2.6 Approx 0.8<br />

Granite 1450 0.79 1.9 Approx 0.85<br />

50% Ethylene Glycol; 50% Water 0 3.5 3.7 Approx 0.3<br />

Water 0 4.2 4.2 0.34<br />

The specific energies of present-day<br />

thermal storage systems lie in the range<br />

0.032 – 0.036 MJ/kg, comparable with<br />

batteries but at three orders of magnitudes<br />

smaller than hydrocarbon fuel. What is the<br />

upper limit on their performance To<br />

answer that we need the thermal<br />

properties of thermal storage materials.<br />

Table 5 lists some of these.<br />

The element with the greatest specific<br />

heat capacity is lithium (3.6 kJ/kg/K).<br />

Lithium melts at 180°C, but can be stored<br />

in its liquid state, conceivably up to<br />

1000°C. If we assume that the specific<br />

heat capacity of liquid lithium is the same<br />

as that of the solid and add its latent heat<br />

of fusion (~0.4 MJ/kg), the total energy<br />

density becomes ~3.5 MJ/kg.<br />

Latent heat storage offers substantial<br />

specific energy, stored and recovered over<br />

a narrow temperature range. The melting<br />

point determines the output temperature,<br />

so must be chosen to match the<br />

application. The energy density is then<br />

determined by the latent heat of fusion<br />

(Table 5).<br />

Thermal energy storage creates no<br />

alarming demands for critical materials.<br />

Meeting the scenario described earlier<br />

could require 0.6% of the world production<br />

of chromium and 5% of that of nickel.<br />

8. Batteries<br />

Batteries store chemical energy. There<br />

two classes of battery. Primary batteries<br />

can only be used once, after which they<br />

can be recycled, but not recharged.<br />

Secondary batteries can be recharged<br />

100-10,000+ times before they fail. Here,<br />

we consider six types of secondary<br />

batteries which have shown promise in<br />

utility scale or automotive scale energy<br />

storage: lead-acid, lithium ion (Li-ion),<br />

sodium-sulfur (NaS), nickel-cadmium<br />

(Ni-Cd or NiCad), nickel-metal hydride<br />

(Ni-MH), and vanadium redox flow<br />

batteries (VRB). Appendix 2 contains bills<br />

of material for each.<br />

A battery consists of two half cells, each<br />

containing a metal and a salt solution of<br />

that metal (e.g., metal sulfate). The half<br />

cell with the more reactive metal (metal A,<br />

in this example) is the anode. The metal is<br />

oxidized:<br />

A →<br />

+ 2x<br />

A<br />

+<br />

−<br />

2 x e<br />

Materials for Energy Storage Systems 18 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Figure 17. Left: diagram of the discharge of a lead-acid battery.<br />

Right: Demand ratios for antimony and lead used in lead acid batteries.<br />

*These are not critical, but large scale implementation is expected to stretch their supply.<br />

Figure 18. Left: Diagram of the discharging of a lithium-ion battery.<br />

Right: Demand ratios for critical materials used in lithium-ion batteries.<br />

The half cell with the less reactive metal<br />

(metal B) is the cathode. The metal ions<br />

from the solution are reduced:<br />

+ 2<br />

B x −<br />

+ 2 x e →<br />

The metals from each half cell are<br />

connected to an external load through a<br />

circuit that transports the electrons from<br />

the anode to the cathode, providing<br />

electrical power. In order to maintain<br />

balance of charge, anions (negatively<br />

charged ions) are allowed to pass through<br />

B<br />

a porous disk from one solution to the<br />

other.<br />

Lead acid batteries (Figure 17) have a<br />

lead anode and a lead dioxide cathode.<br />

On discharge, both of these become lead<br />

sulfate. At the anode:<br />

−<br />

+ −<br />

Pb + HSO4<br />

→ PbSO4<br />

+ H + 2 e<br />

At the cathode the reaction is:<br />

+ − −<br />

2 + H + HSO 4 + 2 e → PbSO 4 2<br />

PbO 3 + H 2 O<br />

Materials for Energy Storage Systems 19 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Figure 19. Left: Diagram of the discharge of a sodium-sulfur battery.<br />

Right: Demand ratios for critical materials used in sodium-sulfur batteries.<br />

Lead-acid batteries do not use critical<br />

materials in significant quantities.<br />

However, large scale implementation will<br />

require an increase in the supply of both<br />

lead and antimony (Figure 17, right).<br />

Li-ion batteries (Figure 18) have an<br />

anode of graphite intercalated with lithium<br />

and a cathode of lithium compounds.<br />

During discharge, lithium ions move from<br />

the graphite anode:<br />

+ −<br />

Li x C6 → x Li + x e + 6 C<br />

The Li + ions diffuse through the separator<br />

and are taken up by the lithium<br />

compounds (typically, a lithium metal<br />

oxide compound, LiMO) at the cathode:<br />

+ −<br />

Li1−<br />

x MO + x Li + x e → LiMO<br />

Large scale adoption of Li-ion for<br />

automotive propulsion or utility electricity<br />

storage may be constrained by the supply<br />

of lithium and graphite. Figure 18, right,<br />

shows that large scale implementation<br />

would result in lithium usage which is<br />

about eighty times today’s supply. The<br />

demand for graphite could exceed three<br />

times today’s supply production.<br />

Sodium-sulfur batteries (Figure 19) have<br />

an anode of molten sodium and a cathode<br />

of molten sulfur. On discharge, the sodium<br />

is oxidized and the ions pass through an<br />

alumina electrolyte to reduce the sulfur<br />

and create sodium polysulfide. The overall<br />

reaction is:<br />

2 Na + 4 S → Na2S4<br />

Graphite is used to contain the molten<br />

sulfur and prevent it from corroding the<br />

current collectors and casing. Large scale<br />

implementation of NaS batteries may<br />

stretch the supply of graphite, as it will use<br />

40% of the annual supply (Figure 19,<br />

right).<br />

Ni-Cd batteries (Figure 20) have a<br />

cadmium-plated anode and a nickel oxidehydroxide<br />

plated cathode. On discharge,<br />

the cadmium at the anode is oxidized:<br />

−<br />

−<br />

Cd + 2 OH → Cd(<br />

OH ) 2 + 2 e<br />

The nickel oxide-hydroxide is reduced:<br />

−<br />

−<br />

NiO(<br />

OH ) + H 2 O + 2 e → Ni(<br />

OH ) 2 + OH<br />

Materials for Energy Storage Systems 20 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Figure 20. Left: Diagram of the discharge of a nickel-cadmium battery. Right: Demand ratios for critical<br />

materials used in nickel-cadmium batteries. *Cadmium is included because, even though it is not a critical<br />

material, large scale implementation is likely to stretch its supply.<br />

Figure 21. Left: Diagram of the discharge of a nickel-metal hydride battery.<br />

Right: Demand ratios for critical materials used in nickel-metal hydride batteries.<br />

Large scale implementation of Ni-Cd<br />

batteries could be constrained by of<br />

supply of nickel, cobalt, and lithium<br />

(Figure 20, right). Although cadmium is not<br />

classified as a critical material, assuring its<br />

supply could be a problem.<br />

Nickel-metal hydride batteries<br />

(Figure 21) resemble Ni-Cd batteries, but<br />

the anode is plated with a metal hydride<br />

rather than cadmium. On discharge, the<br />

metal hydride is oxidized:<br />

−<br />

−<br />

MH + OH → M + H 2O<br />

+ 2 e<br />

The nickel is reduced, as in the Ni-Cd<br />

batteries:<br />

−<br />

−<br />

NiO(<br />

OH ) + H 2 O + 2 e → Ni(<br />

OH ) 2 + OH<br />

Materials for Energy Storage Systems 21 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Figure 22. Left: Diagram of the discharge of a vanadium redox flow battery. Right: Demand ratios for critical<br />

materials used in vanadium flow batteries. *Vanadium is included because, although it is not a critical<br />

material, large scale implementation is likely to stretch its supply.<br />

The anode, typically, is a hydride of<br />

lanthanum, neodymium, praseodymium,<br />

or cerium. Of these, cerium has the largest<br />

annual production, so we assume that the<br />

large scale implementations would use<br />

cerium hydride. This would result in a<br />

bottleneck of cerium and nickel supply<br />

(Figure 21, right). Even if the large scale<br />

implementation used a mixture of the<br />

available metals, such that only ¼ of the<br />

cerium was used, this would still result in<br />

the use of 200 times the current annual<br />

production—and this ratio would be larger<br />

for the other elements.<br />

Vanadium redox batteries (Figure 22)<br />

differ from other batteries. The electrolytes<br />

are stored in tanks, and are pumped<br />

through the battery cell, where they are<br />

reduced or oxidized. The anode and<br />

cathode electrolytes are made up of a<br />

solution of vanadium in different oxidation<br />

states. The anode electrolyte is a solution<br />

of vanadium (II) ions, which are oxidized<br />

to vanadium (III) ions on discharge:<br />

+ 2<br />

V<br />

+ 3<br />

→ V<br />

−<br />

+ e<br />

The cathode electrolyte is a solution of<br />

vanadium (V) oxide ions, which are<br />

reduced to vanadium (IV) oxide ions:<br />

+ + − + 2 −<br />

VO2<br />

+ H + e → VO + OH<br />

Large scale implementation of VRBs<br />

would use a significant fraction of today’s<br />

production of graphite (Figure 22, right)<br />

and would stretch the supply of vanadium,<br />

even though this is not a critical material.<br />

The ultimate limits for batteries<br />

To determine the maximum theoretical<br />

specific energy of batteries, we need to<br />

consider the standard electrode potential<br />

(SEP) of various half-cells. Consider, as<br />

an example, a lithium anode and a fluorine<br />

cathode (ignoring any practicalities of<br />

making this safe!), a combination with an<br />

exceptional high SEP differences. The<br />

oxidation of lithium<br />

+ −<br />

Li → Li + e<br />

has a standard electrode potential of<br />

3.04 V, and the reduction of fluorine<br />

F + 2 e<br />

− → 2 F<br />

−<br />

2<br />

has a standard electrode potential of<br />

2.87 V. The overall reaction<br />

2 Li + F2 → 2 LiF<br />

Materials for Energy Storage Systems 22 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


gives a cell voltage of V=5.91 V. For one<br />

mole of reactants, having a mass of<br />

51.9 g, the two moles of electrons<br />

exchanged, with charge C = 2 x 96,485<br />

Coulombs, will deliver the energy:<br />

W = V C = 5 .91 x 2 96,485 = 1.14 MJ / mole<br />

corresponding to a specific energy of<br />

22 MJ/kg, comparable with that of<br />

conventional fuels (20 – 50 MJ/kg). Safety<br />

concerns have held back the development<br />

of the lithium-fluorine cell. The closest<br />

technology is the lithium-carbon<br />

monofluoride (Li-CFx) cell, which has the<br />

combined reaction:<br />

x Li + ( CF)<br />

x → x LiF +<br />

x C<br />

The carbon maintains the fluorine in a safe<br />

compound, but reduces the theoretical<br />

energy density to 7.9 MJ/kg, which we will<br />

take as an upper limit for battery<br />

technology. In practice real (Li-CFx) cells<br />

have only achieved energy densities in the<br />

range 0.94 – 2.8 MJ/kg, and they are not<br />

rechargeable.<br />

9. Hydrogen Energy Storage<br />

Energy can be stored by using it to<br />

generate hydrogen by electrolysis of water<br />

(Figure 23). Energy is recovered by<br />

passing the hydrogen to a fuel cell to<br />

generate electricity. Electrolysis breaks<br />

the water into positive H + ions (or protons)<br />

and oxygen. At the anode:<br />

+ −<br />

2 H 2O<br />

→ O2<br />

+ 4 H + 4 e<br />

At the cathode:<br />

+ −<br />

2 H + 2 e → H 2<br />

In the fuel cell hydrogen passes the<br />

anode, where it is oxidized to hydrogen<br />

ions (or protons) and electrons at the<br />

anode:<br />

+ −<br />

H 2 → 2 H + 2 e<br />

At the cathode, oxygen is reduced with the<br />

hydrogen to form water:<br />

+ −<br />

O2 + 4 H + 4 e → 2 H 2O<br />

The electrolyte allows passage of<br />

hydrogen ions but not electrons. The<br />

electrons flow through an external circuit,<br />

Figure 23. Diagram of hydrogen energy storage.<br />

Materials for Energy Storage Systems 23 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


where they deliver energy to a load.<br />

Catalysts are required at the cathode and<br />

anode, platinum at the anode, and nickel<br />

at the cathode. Their use raises concerns<br />

over supply: large scale deployment might<br />

use ~25% of nickel supply and 400% of<br />

world production of platinum (Figure 24).<br />

Figure 24. Demand ratios for critical materials used<br />

in hydrogen storage systems.<br />

The performance metrics of hydrogen<br />

storage presented in Table 2, and used in<br />

subsequent charts, are based on<br />

conceptual designs of a hydrogen storage<br />

system for which a number of<br />

assumptions have been made. The data is<br />

based on hydrogen generators and fuel<br />

cells (Hydrogenics, 2010). The mass of<br />

the hydrogen generator was not included<br />

in the onboard hydrogen storage. The<br />

specific energy and energy density of the<br />

stored hydrogen are based on the US<br />

Department of Energy 2010 targets for<br />

hydrogen storage of 5.4 MJ/kg and<br />

3.2 MJ/m 3 (US Department of Energy,<br />

2011), which have been met by MOF-177<br />

(a zinc compound that absorbs hydrogen<br />

as a hydride).<br />

The US Department of Energy has also<br />

set ultimate targets for the specific energy<br />

and energy density of hydrogen storage:<br />

9 MJ/kg and 8,300 MJ/m 3 . These can be<br />

taken as the practical limits on hydrogen<br />

energy storage, assuming that they will<br />

not be significantly affected by the mass of<br />

the fuel cell and hydrogen generator.<br />

These are barely competitive with<br />

conventional fuels, but many automotive<br />

companies are already investing heavily in<br />

research on hydrogen power vehicles.<br />

10. Capacitors and supercapacitors<br />

A capacitor stores energy as electrical<br />

charge. At its simplest it consists of two<br />

closely-spaced plates with a dielectric<br />

between them (Figure 25, left). Energy is<br />

used to move charge from one plate to the<br />

other, creating a potential difference<br />

between them. Energy is extracted by<br />

allowing the charge to return, passing<br />

through an external load. Capacitors can<br />

be charged and discharged very quickly<br />

giving high specific power, but their<br />

specific energy is small, about<br />

0.0001 MJ/kg.<br />

Figure 25. Left A capacitor. Right: A super-capacitor<br />

Materials for Energy Storage Systems 24 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


The energy stored in a capacitor is<br />

1 2<br />

W = C V<br />

2<br />

where C is the capacitance and V is the<br />

voltage. However, the electric field cannot<br />

exceed the breakdown field, V d , which<br />

limits the voltage to V = Vd<br />

d , where d is<br />

the distance between plates. The<br />

capacitance of a capacitor is given by<br />

C = ε A / d where ε = ε r ε o is the electric<br />

permittivity of the dielectric, ε r is the<br />

relative permittivity, ε o is the permittivity of<br />

free space, and A is the area of the<br />

plates. Assembling these results gives the<br />

maximum energy that the capacitor can<br />

contain without breakdown:<br />

1<br />

2<br />

Wmax<br />

= ε r ε o A d V<br />

2<br />

d<br />

and a maximum specific energy<br />

Wmax<br />

m<br />

2<br />

1 Vd<br />

= ε r ε o<br />

2 ρ<br />

where m is the mass and ρ is the<br />

density of the dielectric. The maximum<br />

specific energy of a number of dielectrics,<br />

listed in Table 6, suggests that capacitors<br />

with a styrene-butadiene dielectric might<br />

double the performance to 0.0002 MJ/kg.<br />

Super-capacitors offer greater<br />

performance. Super-capacitors store the<br />

charges at the interface between activated<br />

carbon and a liquid electrolyte (Figure 25,<br />

right), rather than between two plates.<br />

Because of the large surface to volume<br />

ratio of activated carbon, and the<br />

vanishingly thin distance over which the<br />

charge is stored, super-capacitors have<br />

greater capacitance and energy densities<br />

of 0.01-0.1 MJ/kg. The band-gap of<br />

dielectric materials limits the maximum<br />

energy density achievable in supercapacitors<br />

to about 2 MJ/kg (House,<br />

2009), still an order of magnitude less than<br />

that of conventional fuels.<br />

Super-capacitors do not use critical<br />

materials so that large scale<br />

implementation would not result in supply<br />

bottlenecks.<br />

Dielectric<br />

Table 6. Properties of dielectric materials 5 .<br />

Dielectric Breakdown Density, ρ<br />

constant, ε r field, V d (kg/m 3 )<br />

(MV/m)<br />

Specific energy<br />

(MJ/kg)<br />

Syrene-Butadiene Block<br />

Copolymer<br />

2.45-2.55 85-140 1,000 0.0001-0.0002<br />

Cyclo Olefin Polymer 2.13-2.47 67-73 1,000 0.00005<br />

Transparent Polyamide<br />

(nylon)<br />

3.3-4.6 49-51 1,000 0.00004-0.00005<br />

Vermiculite 6-8 8-14 64-160 0.00002-0.00007<br />

Mica 5.4-8.7 40-79 2,600-3,200 0.00001-0.00007<br />

Alumina (97.6) 9.0-9.5 41-45 3,700-3,800 0.00002<br />

5 Data from CES EduPack<br />

Materials for Energy Storage Systems 25 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


11. Superconducting Magnetic<br />

Energy Storage (SMES)<br />

Superconducting magnetic energy storage<br />

(SMES) systems store energy as a<br />

magnetic field, set up by passing current<br />

through a superconducting coil. Because<br />

there is no resistance, the current and the<br />

field, once established, continues to flow<br />

without consuming further power. Energy<br />

is recovered by discharging the current<br />

through an external load.<br />

Superconducting magnets require<br />

refrigeration and that does consume<br />

power, reducing efficiency, especially over<br />

large periods. High temperature<br />

superconductors (HTS) such as YBCO—<br />

yttrium barium copper oxide—which<br />

become superconducting at ~77K require<br />

less power for refrigeration than low<br />

temperature ~4K superconductors (LTS),<br />

but they still need some. For this reason,<br />

SMES tends to be used for short term<br />

storage (e.g., for frequency regulation).<br />

The energy density stored in a magnetic<br />

field is:<br />

Based on current HTS-SMES systems,<br />

large scale implementation would result in<br />

a bottleneck of supply of Yttrium, essential<br />

for the superconducting YBCO magnets<br />

(Figure 26).<br />

Figure 26. Demand ratios for critical materials<br />

used in superconducting magnetic energy<br />

storage.<br />

W<br />

V<br />

=<br />

2<br />

1 B<br />

2 µ<br />

o<br />

where B is the magnetic flux density and<br />

µ o is the permittivity of free space. The<br />

energy density is maximized by<br />

maximizing B , which, for a type-II<br />

superconductor, cannot exceed the upper<br />

critical field of the magnet windings. The<br />

upper critical field of YBCO is 250 T at a<br />

temperature of zero Kelvin. If 250 T can<br />

be achieved, an energy density of<br />

25,000 MJ/m 3 can be achieved. Assuming<br />

a similar physical density to today’s SMES<br />

devices, this would result in a specific<br />

energy of 50 MJ/kg. This is competitive<br />

with fuels, but is only achieved at<br />

unrealistically low temperatures, and does<br />

not include the mass of the cryostat and<br />

ancillary devices.<br />

Materials for Energy Storage Systems 26 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Appendix 1 Definition of terms<br />

Resource intensities of<br />

construction<br />

Capital intensity<br />

US$/MJ<br />

The capital intensity of an energy-storage<br />

system is the cost of building or<br />

purchasing the system per megajoule of<br />

storage capacity. As an example, a typical<br />

alkaline AA battery has a capacity of<br />

around 13.5x10 -3 MJ and costs about a<br />

dollar. Its capital intensity is:<br />

Capital intensity<br />

= Cost/Storage capacity<br />

= 1 / 0.0135<br />

= 74 $/MJ<br />

Area intensity m 2 /MJ<br />

The area intensity of an energy storage<br />

system is the area of land that the system<br />

typically occupies per megajoule of<br />

storage capacity. As an example, the<br />

Ffestiniog pumped hydro power station<br />

has an area of 340,000 m² and can store<br />

4.7 x 10 6 MJ. Its area intensity is:<br />

Area intensity<br />

= Area/Storage capacity<br />

= 340,000 / (4.7 x 10 6 )<br />

= 0.07 m 2 /MJ<br />

Material intensity<br />

kg/MJ<br />

The material intensity of an energy<br />

storage system is the mass of material<br />

that the system typically requires per<br />

megajoule of storage capacity. As an<br />

example, a particular Bosch car battery<br />

has a capacity of 4.36 MJ and a mass of<br />

23.2 kg. Its mass intensity is calculated as<br />

Material intensity<br />

= Mass of material/Storage capacity<br />

= 23.2 / 4.36<br />

= 5.4 kg/MJ<br />

The bill of materials for a given storage<br />

system allows the material intensity to be<br />

broken down by material. These materialspecific<br />

intensities can then be compared<br />

to the annual world production of that<br />

material to flag up systems that, if<br />

deployed widely, might be constrained by<br />

material supply.<br />

Energy intensity<br />

MJ/MJ<br />

The energy intensity of an energy storage<br />

system is the energy required to create<br />

the system per megajoule of storage<br />

capacity. This includes the energy needed<br />

to extract and process raw materials,<br />

fabricate components, and construct the<br />

final plant, with transport at different<br />

stages of the construction process taken<br />

into account. As an example, the energy<br />

needed to build a hypothetical molten salt<br />

thermal storage system with a capacity of<br />

7.2x10 6 MJ was calculated at 9.2x10 8 MJ.<br />

Its energy intensity is:<br />

Energy intensity<br />

= Construction energy/Storage capacity<br />

= 9.2 x 10 8 /(7.2 x 10 6 )<br />

= 128 MJ/MJ<br />

CO2<br />

intensity<br />

kg/MJ<br />

The CO 2 intensity (or carbon intensity) of<br />

an energy storage system is the CO 2<br />

(equivalent) released to the atmosphere<br />

as a consequence of its construction the<br />

system per megajoule of storage capacity.<br />

As an example, a hypothetical thermal<br />

energy storage system has a capacity of<br />

7.2x10 6 MJ and causes the emission of<br />

6.4x10 7 kg of CO 2 . The carbon intensity is:<br />

CO 2 intensity<br />

= CO 2 emission/Storage capacity<br />

= 6.4 x 10 7 / (7.2 x 10 6 )<br />

= 8.9 kg/MJ<br />

Operational parameters<br />

Specific energy<br />

MJ/kg<br />

The specific energy of an energy-storage<br />

system is the energy the system can store<br />

per unit of its mass. As and example, a<br />

silver oxide button cell has a mass of 0.7 g<br />

(7x10 -4 kg) whilst storing 2.2x10 -4 MJ. Its<br />

specific energy is<br />

Specific energy<br />

= Energy stored/Mass<br />

= 2.2 x 10 -4 / (7 x 10 -4 )<br />

= 0.31 MJ/kg<br />

Energy density<br />

MJ/m<br />

The energy density of an energy-storage<br />

system is the energy the system can store<br />

per unit of volume. As an example, a lead-<br />

3<br />

Materials for Energy Storage Systems 27 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


acid battery measuring 120x80x80 mm<br />

can store 0.3 MJ. Its energy density is:<br />

Energy density<br />

= Energy stored/Volume<br />

= 0.3 / (7.7 x 10 -4 )<br />

3<br />

= 391 MJ/m<br />

Specific power<br />

W/kg<br />

The specific power is the rate at which<br />

energy can be drawn from the system per<br />

unit of its mass. As an example, the<br />

Maxwell VCAP350 super-capacitor weighs<br />

63 g (6.3x10 -2 kg) and can discharge with<br />

a continuous power of 67.5 W. Its specific<br />

power is calculated to be<br />

Specific power<br />

= Energy per second/Mass<br />

= 67.5 / 0.063<br />

= 1071 W/kg<br />

Economic energy-storage capacity<br />

W/kg<br />

The economic energy storage capacity is<br />

the range of energy for which a particular<br />

storage system is economically viable. As<br />

an example, it is impractical to create a<br />

pumped hydro storage plant for a tiny<br />

amount of amounts of energy. It is equally<br />

impractical to use silver oxide batteries to<br />

store megajoules of energy.<br />

Economic energy power capacity W/kg<br />

The economic power capacity of an<br />

energy storage technology is the range of<br />

power outputs that systems can deliver<br />

economically.<br />

Cycle efficiency %<br />

The cycle efficiency is the percentage of<br />

the energy put into a system that is<br />

recovered when the energy is retrieved.<br />

This is measured for a typical cycle time,<br />

and will usually decrease if cycle is<br />

unusually long. As examples, a battery will<br />

slowly discharge if unused and frictional<br />

losses of a flywheel increase the longer it<br />

is spinning.<br />

Cycle efficiency<br />

= (Energy output / Energy input) x 100<br />

Cycle life -<br />

The cycle life is the number of times an<br />

energy storage system can charged and<br />

discharged before the capacity of the<br />

system drops below 80% of its initial<br />

capacity. Cycle life is limited by factors<br />

such as fatigue in mechanical systems<br />

and electrolyte degradation in<br />

electrochemical systems.<br />

Operating cost<br />

US$/MJ<br />

The operating cost is the approximate cost<br />

of one charge/discharge cycle of one unit<br />

of energy, and includes the cost of<br />

maintenance, heating, labor, etc.<br />

Adaptable for mobile systems Yes/No<br />

Adaptable for mobile systems. If ticked<br />

(meaning Yes) the energy storage system<br />

is suitable for providing mobile power such<br />

as powering an automobile or a portable<br />

communication device.<br />

Upper limits for performance<br />

metrics<br />

Theoretical max specific energy MJ/kg<br />

The maximum specific energy of an<br />

energy storage system is the theoretical<br />

upper limit to the energy it can store per<br />

unit mass. As an example, the theoretical<br />

strength, roughly E/10 (where E is<br />

Young’s modulus) sets an absolute upper<br />

limit to the energy that could be stored in<br />

springs and flywheels.<br />

3<br />

Theoretical max energy density MJ/m<br />

The maximum energy density of an<br />

energy storage system is the theoretical<br />

upper limit to the energy it can store per<br />

unit volume. As an example, the<br />

theoretical strength, roughly E/10 (where<br />

E is Young’s modulus) sets an absolute<br />

upper limit to the energy that could be<br />

stored in springs and flywheels.<br />

Status<br />

Current installed capacity<br />

GJ<br />

The current installed capacity is the sum<br />

of the energy storage capacities of all<br />

systems using a certain technology<br />

worldwide.<br />

Growth rate<br />

% per year<br />

The growth rate is the rate of additional<br />

energy capacity added per year,<br />

expressed as a percentage of the current<br />

installed capacity.<br />

Materials for Energy Storage Systems 28 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Appendix 2: Approximate<br />

Material Intensities for Energy<br />

Storage Systems<br />

Bills of materials for energy storage<br />

systems are assembled here. They are<br />

expressed as material intensities, I m ,<br />

meaning mass (kg) of each material per<br />

unit (MJ) of energy storage capacity.<br />

There is no systematic, self consistent<br />

assembly of such data of which we are<br />

aware, so it has to be patched together<br />

from diverse sources. These differ in detail<br />

and scope. Some, for instance, are limited<br />

to the system alone, others include the<br />

copper and other materials needed to<br />

connect the system to the grid. Others<br />

give indirect information from which<br />

missing material content can be inferred.<br />

So be prepared for inconsistencies.<br />

Despite these difficulties there is enough<br />

information here to draw conclusions<br />

about the demand that a commitment to<br />

any one of them would put on material<br />

supply. The resource-demand plots in the<br />

text use the data in these tables. They are<br />

based on an imagined scenario: that in<br />

order to meet global electric power<br />

demand, the capacity of a chosen power<br />

system must be expanded by 200 GW per<br />

Pumped Hydro Storage<br />

year (See the Materials for Low Carbon<br />

Power White Paper), and to compliment<br />

this, 10 hours of storage capacity must be<br />

built—meaning 2,000 GW.hr or 7.2 PJ<br />

(equal to 7.2 x 10 6 MJ) of storage capacity<br />

per year. The metric for resource pressure<br />

Rs<br />

is the mass of each material required<br />

for the expansion of the system per year,<br />

6<br />

7.2 x10<br />

I m , expressed in kg/year divided<br />

by the current (2010) global production per<br />

year, , also expressed in kg/year.<br />

P a<br />

6<br />

Rs<br />

= 7.2 x10<br />

I m<br />

Pa<br />

The plots show this as the percent<br />

demand of current production.<br />

The resource demand is not interesting<br />

when it is trivial. The plots show the<br />

demand on the materials that are deemed<br />

to be “critical”, as explained in Section 2,<br />

together with other (non-critical) materials<br />

whose supply will likely be stretch by large<br />

scale implementation of the particular<br />

system. Their global annual productions<br />

for critical materials in 2010 appear Figure<br />

10. They are marked with an asterisk (*) in<br />

the tables below.<br />

Material<br />

Intensity (kg/MJ)<br />

Aluminum 0.01<br />

Brass 0.0002-0.0014<br />

Chromium* 0.005-0.01<br />

Concrete 60-120<br />

Copper 0.0014-0.0049<br />

Iron 0.60-0.79<br />

Lead 0.0006-0.005<br />

Magnesium 0.0002-0.002<br />

Manganese* 0.0004-0.003<br />

Molybdenum 0.0005-0.004<br />

Plastics 0.012-0.02<br />

Wood 0.16-1.3<br />

Zinc 0.0008-0.01<br />

Total mass, all materials 61-120<br />

Materials and quantities (Tahara et al., 1997), (Pacca & Horvath, 2002), (Vattenfall AB<br />

Generation Nordic, 2008) and (Ribeiro & da Silva, 2010).<br />

Materials for Energy Storage Systems 29 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Compressed Air Energy Storage<br />

Material<br />

Intensity (kg/MJ)<br />

Carbon steel 0.0084-0.054<br />

Chromium* 0.000006-0.00013<br />

Concrete 1.8-12<br />

Copper 0.00025-0.0016<br />

High Alloy Steel 0.086-0.56<br />

Iron 0.0045-0.029<br />

Manganese* 0.0011-0.0069<br />

Molybdenum 0.000011-0.000069<br />

Plastics 0.00093-0.0060<br />

Silicon 0.00024-0.0015<br />

Vanadium 0.000032-0.00021<br />

Total mass, all materials 1.9-13<br />

Materials and quantities from (Meier & Kulcinski, 2000).<br />

Flywheels<br />

Material<br />

Intensity (kg/MJ)<br />

Carbon Fiber 1.8<br />

Carbon Steel 12-400<br />

Copper 1-30<br />

Plastics 2-4<br />

YBCO 0.0058-0.17<br />

of which Barium 0.0024-0.07<br />

of which Copper 0.0017-0.051<br />

of which Yttrium* 0.00078-0.023<br />

Total mass, all materials 17-500<br />

Materials and quantities from (Hartikainen et al., 2007).<br />

Thermal Storage<br />

Material<br />

Intensity (kg/MJ)<br />

Calcium Silicate 0.06-0.12<br />

Carbon Steel 0.78-1.4<br />

Chromium* 0.008-0.03<br />

Concrete 3.1-4.8<br />

Foam Glass 0.041-0.084<br />

Mineral Wool 0.15-0.26<br />

Molten Salt<br />

7.1-24<br />

(40% KNO 3 , 60% NaNO 3 )<br />

Nickel* 0.0044-0.017<br />

Nitrogen 0.026-0.4<br />

Refractory Brick 0.40-0.62<br />

Silica 0-17<br />

Total mass, all materials 12-48<br />

Materials and quantities from (Heath et al., 2009).<br />

Materials for Energy Storage Systems 30 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Lead Acid Batteries<br />

Material<br />

Intensity (kg/MJ)<br />

Antimony 0.055-0.14<br />

Copper 0.017-0.042<br />

Glass 0.11-0.28<br />

Lead 1.4-3.5<br />

Lead oxides 1.9-4.9<br />

Plastics 0.55-1.4<br />

Sulfuric Acid 0.55-1.4<br />

Total mass, all materials 4.6-12<br />

Materials and quantities from (Sullivan & Gaines, 2010).<br />

Lithium-Ion Batteries<br />

Material<br />

Intensity (kg/MJ)<br />

Aluminum 0.072<br />

Carbon Black 0.075-0.09<br />

Copper 0.13<br />

Ethylene Glycol Dimethyl Ether 0.44<br />

Graphite* 0.47<br />

Lithium Compounds 1.2<br />

of which Lithium* 0.2<br />

Plastics 0.18-0.2<br />

Total mass, all materials 1.5-2.6<br />

Based on Iron Phosphate type Lithium Ion Batteries. Materials and quantities from (Zackrisson et al., 2010).<br />

Sodium-Sulfur Batteries<br />

Material<br />

Intensity (kg/MJ)<br />

Alumina 0.18-0.63<br />

Aluminum 0.32-1.1<br />

Carbon Steel 0.18-0.64<br />

Copper 0.05-0.17<br />

Glass 0.06-0.22<br />

Graphite* 0.03-0.1<br />

Plastics 0.11-0.4<br />

Silica 0.21-0.76<br />

Sodium 0.11-0.4<br />

Sulfur 0.18-0.63<br />

Total mass, all materials 1.4-5<br />

Materials and quantities from (Sullivan & Gaines, 2010).<br />

Materials for Energy Storage Systems 31 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Nickel-Cadmium Batteries<br />

Material<br />

Intensity (kg/MJ)<br />

Cadmium 1.1-3.1<br />

Carbon Steel 0.38-1.1<br />

Chromium* 0.091-0.26<br />

Cobalt* 0.06-0.18<br />

Copper 0.18-0.51<br />

Lithium Hydroxide 0.03-0.09<br />

of which Lithium* 0.0087-0.025<br />

Nickel and Compounds 1.6-4.7<br />

of which Nickel* 1.3-3.9<br />

Plastics 0.13-0.39<br />

Potassium Hydroxide 0.22-0.65<br />

Total mass, all materials 3.8-11<br />

Materials and quantities from (Sullivan & Gaines, 2010) and (Gaines & Singh, 1995).<br />

Nickel-Metal Hydride Batteries<br />

Material<br />

Intensity (kg/MJ)<br />

Carbon Steel 1.2-3.5<br />

Cerium* 0.22-0.64<br />

Nickel and Compounds 0.81-2.4<br />

of which Nickel* 0.66-2.0<br />

Plastics 0.14-0.04<br />

Potassium Hydroxide 0.08-0.24<br />

Total mass, all materials 2.5-6.8<br />

Materials and quantities from (Sullivan & Gaines, 2010).<br />

Vanadium Flow Batteries<br />

Material<br />

Intensity (kg/MJ)<br />

Carbon Black 0.037-0.065<br />

Carbon Steel 0.85-1.5<br />

Copper 0.064-0.11<br />

Graphite* 0.024-0.042<br />

Plastics 0.31-0.54<br />

Sulfuric Acid 2.1-3.6<br />

Vanadium 0.45-0.78<br />

Total mass, all materials 3.8-6.6<br />

Materials and quantities from (Rydh, 1999).<br />

Materials for Energy Storage Systems 32 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Hydrogen Storage<br />

Material<br />

Intensity (kg/MJ)<br />

Aluminum 0.03<br />

Carbon Black 0.01<br />

Carbon Steel 1.7<br />

Chromium* 0.09<br />

Copper 0.2<br />

Glass Fiber 0.06<br />

Graphite* 0.24<br />

Molybdenum 0.0007<br />

Nickel* 0.05<br />

Phosphoric Acid 0.02<br />

Plastics 0.04<br />

Platinum* 0.000003<br />

Silicon Carbonate 0.12<br />

Zinc 0.06<br />

Total mass, all materials 2.6<br />

Materials and quantities from (van Rooigen, 2006) and (Nadal, 1997).<br />

Super-capacitors<br />

Material<br />

Intensity (kg/MJ)<br />

Activated Carbon 12-14<br />

Aluminum 9.8-20<br />

Plastics 1-2.5<br />

Ammonium Salts in Acetonitrile 23-29<br />

Total mass, all materials 45-66<br />

Materials and quantities from (Maxwell Technologies, Inc., 2009) and (Maxwell Technologies, Inc., 2011).<br />

Superconducting Magnetic Energy Storage<br />

Material<br />

Intensity (kg/MJ)<br />

Carbon Steel 32<br />

Copper 3.8<br />

Sulfuric Acid 0.22<br />

YBCO 0.52<br />

of which Barium 0.21<br />

of which Copper 0.15<br />

of which Yttrium* 0.07<br />

Total mass, all materials 36<br />

Materials and quantities from (Hartikainen et al., 2007).<br />

Materials for Energy Storage Systems 33 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Further Reading<br />

The starting point—sources that help with the big picture<br />

Electricity Storage Association (2009) “Technology Comparison” Electricity Storage<br />

Association, Washington, DC USA<br />

http://www.electricitystorage.org/technology/storage_technologies/technology_comparison<br />

(Accessed 08/11)<br />

Harvey, L.D.D. (2010) “Mitigating the adverse effect of fluctuations in available wind energy”<br />

in “Energy and the New Reality 2: Carbon-Free Energy Supply” Earthscan Publishing,<br />

London UK. ISBN 978-1-84971-073-2 pp. 129-138<br />

House, K.Z. (2009) “The limits of energy storage technology” The Bulletin of Atomic<br />

Sciences http://www.thebulletin.org/web-edition/columnists/kurt-zenz-house/the-limits-ofenergy-storage-technology<br />

(Accessed 07/11)<br />

MacKay D.J.C. (2009) “Fluctuations and Storage” in “Sustainable energy – without the hot<br />

air” UIT Publishers, Cambridge UK. ISBN 978-0-9544529-3-3 and<br />

http://www.withouthotair.com pp. 186-202<br />

Rastler, D. (2009) “Overview of Electric Energy Storage Options for the Electric Enterprise”<br />

Electric Power Research Institute, CA USA<br />

http://www.energy.ca.gov/2009_energypolicy/documents/2009-04-<br />

02_workshop/presentations/0_3%20EPRI%20-%20Energy%20Storage%20Overview%20-<br />

%20Dan%20Rastler.pdf (Accessed 08/11)<br />

Rastler, D. (2011) “Energy Storage Technology Status” Electric Power Research Institute,<br />

CA USA<br />

http://mydocs.epri.com/docs/publicmeetingmaterials/1104/4ANZGWRJTYQ/E236208_02._R<br />

astler_Energy_Storage_Technology_Status.pdf (Accessed 08/11)<br />

Ribeiro, P.F., Johnson, B.K., Crow, M.L., Arsoy, A. and Liu, Y. (2001) “Energy storage<br />

systems for advanced power applications” Proceedings of the IEEE Vol. 89 (12) pp. 1744-<br />

1756<br />

Schaber, C., Mazza, P. and Hammerschlag, R. (2004) “Utility-Scale Storage of Renewable<br />

Energy” The Electricity Journal Vol. 17 (6) pp. 21-24<br />

Schoenung S.M. (2001) “Characteristics and Technologies for Long-vs. Short-Term Energy<br />

Storage” Sandia National Laboratories, CA USA http://prod.sandia.gov/techlib/accesscontrol.cgi/2001/010765.pdf<br />

(Accessed 08/11)<br />

Schoenung, S.M. and Hassenzahl, W.V. (2003) “Long- vs. Short-Term Energy Storage<br />

Technologies Analysis” Sandia National Laboratories, CA USA<br />

http://prod.sandia.gov/techlib/access-control.cgi/2003/032783.pdf (Accessed 08/11)<br />

Pumped Hydro Storage<br />

Denholm, P. and Kulcinski, G.L. (2004) “Life cycle energy requirements and greenhouse gas<br />

emissions from large scale energy storage systems” Energy Conversion and Management<br />

Vol. 45 (13-14) pp. 2153-72<br />

Pacca, S. and Horvath, A. (2002) “Greenhouse Gas Emissions from Building and Operating<br />

Electric Power Plants in the Upper Colorado River Basin” Environmental Science and<br />

Technology, Vol. 36 (14) pp.3194-3200<br />

Ribeiro, F.d.M. and da Silva, G.A. (2010) “Life-cycle inventory for hydroelectric generation: a<br />

Brazilian case study” Journal of Cleaner Production, Vol. 18 (1) pp.44-54<br />

Materials for Energy Storage Systems 34 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Tahara, K., Kojima, T. and Inaba, A. (1997) “Evaluation of CO2 payback time of power<br />

plants by LCA” Proceedings of the Third International Conference on Carbon Dioxide<br />

Removal, Energy Conversion and Management<br />

Vattenfall AB Generation Nordic (2008) “Certified Environmental Product Declaration<br />

EPD(R) of Electricity from Vattenfall's Nordic Hydropower”<br />

http://www.vattenfall.com/en/file/Environmental_Product_Declara_8459904.pdf (Accessed<br />

08/11)<br />

Compressed Air Energy Storage<br />

BINE Informationsdienst (2007) “AA-CAES – Forschungsziele”,<br />

http://www.bine.info/hauptnavigation/publikationen/projektinfos/publikation/druckluftspeicherkraftwerke/aa-caes-forschungsziele/<br />

(Accessed 07/11)<br />

Denholm, P. and Kulcinski, G.L. (2004) “Life cycle energy requirements and greenhouse gas<br />

emissions from large scale energy storage systems” Energy Conversion and Management<br />

Vol. 45 (13-14) pp. 2153-2172<br />

Iowa Stored Energy Park, The (2010) “About The Iowa Stored Energy Park”<br />

http://www.isepa.com/about_isep.asp (Accessed 08/11)<br />

Knight, M. (2008) “Where to store wind-powered energy Under water!” CNN, Technology<br />

http://edition.cnn.com/2008/TECH/science/03/31/windpower/ (Accessed 08/11)<br />

Meier, P.J. and Kulcinski, G.L. (2000) “Life-Cycle Energy Cost and Greenhouse Gas<br />

Emissions for Gas Turbine Power” Energy Centre of Wisconsin<br />

http://www.ecw.org/prod/202-1.pdf (Accessed 08/11)<br />

Messina, J. (2010) “Compressed Air Energy Storage: Renewable Energy” Physorg,<br />

Technology, Energy & Green Tech http://www.physorg.com/news188048601.html<br />

(Accessed 08/11)<br />

Zolfagharifard, E. (2011) “Compressed air energy storage has bags of potential” The<br />

Engineer http://www.theengineer.co.uk/1008374.article (Accessed 08/11)<br />

Springs<br />

Chiu, C.H., Hwan, C.L., Tsai, H.S. and Lee, W.P. (2007) “An experimental investigation into<br />

the mechanical behaviors of helical composite springs” Composite Structures, Vol. 77 (3) pp.<br />

331-340<br />

Hill, F.A., Havel, T.F. and Livermore, C. (2009) “Modeling mechanical energy storage in<br />

springs based on carbon nanotubes” Nanotechnology<br />

Flywheels<br />

Beacon Power “Frequency Regulation and Flywheels”<br />

http://www.beaconpower.com/files/Flywheel_FR-Fact-Sheet.pdf (Accessed 08/11)<br />

Collins, P.G. and Avouris, P. (2000) “Nanotubes for Electronics” Scientific American<br />

Hartikainen, T., Mikkonen, R. and Lehtonen, J. (2007) “Environmental Advantages of<br />

Superconducting Devices in Distrubitued Electricity-Generation” Applied Energy, Vol. 84 (1)<br />

pp. 29-38<br />

EFDA JET (2001) “Flywheel Generators” EDFA JET, Oxfordshire UK<br />

http://www.jet.efda.org/focus-on/jets-flywheels/flywheel-generators/ (Accessed 08/11)<br />

Materials for Energy Storage Systems 35 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Flybrid Systems (2011) “Flybrid Systems Website” http://www.flybridsystems.com/<br />

(Accessed 08/11)<br />

LaMonica, M. (2008) “One megawatt of grid storage, 10 big flywheels” CNET News, Green<br />

Tech http://news.cnet.com/8301-11128_3-9968539-54.html (Accessed 08/11)<br />

Lazarewicz, M. and Judson, J. (2011) “Performance of First 20 MW Commercial Flywheel<br />

Frequency Regulation Plant” Beacon Power Corporation, CA, USA<br />

http://www.beaconpower.com/files/Beacon_Power_presentation_ESA%206_7_11_FINAL.pd<br />

f (Accessed 08/11)<br />

Peng, B. et al. (2008) “Measurements of near-ultimate strength for multiwalled carbon<br />

nanotubes and irradiation-induced crosslinking improvements” Nature Nanotechnology Vol.<br />

3, pp.626-631<br />

Racecar Engineering (2011) “Flywheel hybrid systems (KERS)” http://www.racecarengineering.com/articles/f1/flywheel-hybrid-systems-kers/<br />

(Accessed 08/11)<br />

Thermal Storage<br />

Burkhardt III, J.J., Heath, G.A., Turchi, C.S. (2011) “Life Cycle Assessment of a Parabolic<br />

Trough Concentrating Solar Power Plant and the Impacts of Key <strong>Design</strong> Alternatives”<br />

Environmental Science and Technology Vol. 45 (6) pp. 2457-2464<br />

CALMAC (2007) “Thermal Energy Storage – ICEBANK Model C Specifications and<br />

Drawings” CALMAC Manufacturing Corporation, NJ, USA<br />

http://www.calmac.com/products/icebankc_specs.asp (Accessed 08/11)<br />

The Engineering ToolBox “Heat Storage in Materials”<br />

http://www.engineeringtoolbox.com/sensible-heat-storage-d_1217.html (Accessed 08/11)<br />

Heath, G., Turchi, C., Burkhardt, J., Kutscher, C. and Decker, T. (2009) “Life Cycle<br />

Assessment of Thermal Energy Storage: Two-Tank Indirect and Thermocline” American<br />

Society of Mechanical Engineers (ASME) Third International Conference on Energy<br />

Sustainability, San Francisco<br />

Macnaghten, J. (2010) “Electricity Storage Part 3: Isentropic Electricity Storage” Isentropic<br />

Ltd, Cambridge, UK<br />

http://www.chase.org.uk/files/Isentropic%20ARM%20presentation%20(part%202).pdf<br />

(Accessed 08/11)<br />

Tamme, R. (2009) “Thermal Energy Storage for Large Scale CSP Plants” DLR – Germal<br />

Aerospace Center, Institute of Technical Thermodynamics<br />

http://www.srcf.ucam.org/cuens/img/Academic_Material/Tamme_Storage%20CSP.pdf<br />

(Accessed 08/11)<br />

Batteries<br />

Buchman, I. (2003) “Battery University” http://batteryuniversity.com/ (Accessed 08/11)<br />

Conway, E. (2003) “World’s biggest battery switched on in Alaska” The Telegraph,<br />

Technology http://www.telegraph.co.uk/technology/3312118/Worlds-biggest-batteryswitched-on-in-Alaska.html<br />

(Accessed 08/11)<br />

Denholm, P. and Kulcinski, G.L. (2004) “Life cycle energy requirements and greenhouse gas<br />

emissions from large scale energy storage systems” Energy Conversion and Management<br />

Vol. 45 (13-14) pp. 2153-2172<br />

Energizer “Technical Information” Energizer, MO, USA http://data.energizer.com/ (Accessed<br />

08/11)<br />

Materials for Energy Storage Systems 36 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Gaines, L. and Cuenca, R. (2000) “Costs of Lithium-Ion Batteries for Vehicles” Center for<br />

Transportation Research, Argonne National Laboratory, IL USA<br />

http://www.transportation.anl.gov/pdfs/TA/149.pdf (Accessed 08/11)<br />

Gaines, L. and Nelson, P. (2010) “Lithium-Ion Batteries: Examining Material Demand and<br />

Recycling Issues” Argonne National Laboratory, IL, USA<br />

http://www.transportation.anl.gov/pdfs/B/626.PDF (Accessed 08/11)<br />

Gaines, L. and Singh, M. (1995) “Energy and Environmental Impacts of Electric Vehicle<br />

Battery Production and Recycling” Total Life Cycle Conference and Exposition, Vienna,<br />

Austria<br />

Hartikainen, T., Mikkonen, R. and Lehtonen, J. (2007) “Environmental Advantages of<br />

Superconducting Devices in Distrubitued Electricity-Generation” Applied Energy, Vol. 84 (1)<br />

pp. 29-38<br />

Haruna, H., Itoh, S., Horiba, T., Seki, E. and Kohno, K. (2011) “Large-format lithium-ion<br />

batteries for electrical power storage” Journal of Power Sources Vol. 196 (16) pp. 7002-7005<br />

HILTech Developments Limited (2005) “Vanadium Redox Flow battery – Key Features”<br />

HILTech Developments Limited, Sunderland, UK<br />

http://hiltechdevelopments.com/uploads/news/id43/vanadium_redox.pdf (Accessed 08/11)<br />

Kopera, J.J.C. (2004) “Inside the Nickel Metal Hydride Battery” Cobasys, MI, USA<br />

http://www.cobasys.com/pdf/tutorial/inside_nimh_battery_technology.pdf (Accessed 08/11)<br />

Lotspeich, C. and Van Holde, D. (2002) “Flow Batteries: Has Really Large Scale Battery<br />

Storage Come of Age” Energy Efficiency Center, University of California, CA, USA<br />

NGK Insulators, Ltd. “Principle of the NAS Battery” NGK Insulators, Ltd. Nagoya Japan<br />

http://www.ngk.co.jp/english/products/power/nas/principle/index.html (Accessed 08/11)<br />

Nicad Power (2010) “Technical Data [of Nickel Cadmium Batteries]” Nicad Power Pte Ltd.,<br />

Singapore http://www.nicadpower.com/products/technical/ (Accessed 08/11)<br />

Rydh, C.J. (1999) “Environmental assessment of vanadium redox and lead-acid batteries for<br />

stationary energy storage” Journal of Power Sources Vol. 80 (1-2) pp.21-29<br />

De-Leon, S. (2011) “Li/CFx Batteries, The Renaissance” Schmuel De-Leon Energy, Ltd.,<br />

Israel http://www.contourenergy.com/technology/presentations/sdl_presentation.pdf<br />

(Accessed 08/11)<br />

Skyllas-Kazacos, M. “An Historical Overview of the Vanadium Redox Flow Battery<br />

Development at the University of New South Wales, Autralia” School of Chemical<br />

Engineering & Industrial Chemistry, University of New South Wales, Sydney, Australia<br />

http://www.vrb.unsw.edu.au/overview.htm (Accessed 08/11)<br />

Sullivan, J.L. and Gaines, L. (2010) “A Review of Battery Life-Cycle Analysis: State of<br />

Knowledge and Critical Needs” Energy Systems Division, Argonne National Laboratory, IL,<br />

USA http://www.transportation.anl.gov/pdfs/B/644.PDF (Accessed 08/11)<br />

Wright, R. (2009) “New battery could change world, one house at a time” Daily Herald<br />

http://www.heraldextra.com/news/article_b0372fd8-3f3c-11de-ac77-001cc4c002e0.html<br />

(Accessed 08/11)<br />

Zackrisson, M., Avellán, L. and Orlenius, J. (2010) “Life cycle assessment of lithium-ion<br />

batteries for plug-in hybrid electric vehicles - Critical issues” Journal of Cleaner Production<br />

Vol. 18 (15) pp.1519-1529<br />

Zelinsky, M., Koch, J., Fetcenko, M. (2010) “Heat Tolerant NiMH Batteries for Stationary<br />

Power” Ovonic Battery Company, MI, USA<br />

Materials for Energy Storage Systems 37 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012


Hydrogen Energy Storage<br />

Granovskii, M., Dincer, I., Rosen, M.A. (2006) “Life cycle assessment of hydrogen fuel cell<br />

and gasoline vehicles” International Journal of Hydrogen Energy Vol. 31 (3) pp. 337-352<br />

Hydrogenics (2010a) “Fuel Cell Power Systems based on PEM technology”<br />

http://www.hydrogenics.com/fuel/ (Accessed 08/11).<br />

Hydrogenics (2010b) “Hydrogen Generators based on water electrolysis”<br />

http://www.hydrogenics.com/hydro/ (Accessed 08/11)<br />

Nadal, G. (1997) “Life Cycle Air Emissions from Fuel Cells and Gas Turbines in Power<br />

Generation” London: Imperial College of Science, Technology and Medicine, London UK<br />

http://www3.imperial.ac.uk/portal/pls/portallive/docs/1/7294736.PDF (Accessed 08/11)<br />

US Department of Energy (2011) “DOE Targets for Onboard Hydrogen Storage Systems for<br />

Light-Duty Vehicles”<br />

http://www1.eere.energy.gov/hydrogenandfuelcells/storage/current_technology.html<br />

(Accessed 07/11)<br />

van Rooijen, J. (2006) “A Life Cycle Assessment of the PureCell(TM) Stationary Fuel Cell<br />

System: Providing a Guide for Environmental Improvement” Center for Sustainable Systems,<br />

University of Michigan http://css.snre.umich.edu/css_doc/CSS06-08.pdf (Accessed 07/11)<br />

Capacitors<br />

Kim, I.H., Ma, S.B., Kim, K.B. “Super-capacitor” Lab. Of Energy Conversion & Storage<br />

Materials, Yonsei University, Seoul, Korea<br />

http://web.yonsei.ac.kr/echemlab/public_html/data/sucap.pdf (Accessed 08/11)<br />

Maxwell Technologies, Inc. (2009) “Product Guide - Maxwell Technologies BOOSTCAP<br />

Ultra-capacitors” http://about.maxwell.com/pdf/1014627_BOOSTCAP_Product_Guide.pdf<br />

(Accessed 08/11)<br />

Maxwell Technologies, Inc. (2011) “Production Information Sheet #1004596.3”<br />

http://about.maxwell.com/ultracapacitors/products/pis.asp (Accessed 08/11)<br />

Superconducting Magnetic Energy Storage<br />

ClimateTechWiki “Energy Storage: Superconducting magnetic energy storage (SMES)”<br />

http://climatetechwiki.org/technology/jiqweb-ee (Accessed 08/11)<br />

Hartikainen, T., Mikkonen, R. and Lehtonen, J. (2007) “Environmental Advantages of<br />

Superconducting Devices in Distributed Electricity-Generation” Applied Energy, Vol. 84 (1)<br />

pp. 29-38.<br />

Schnyder, G. and Sjöström, M. (2005) “SMES” École Polytechnique Fédérale de Lausanne,<br />

Switzerland<br />

http://lanoswww.epfl.ch/studinfo/courses/cours_supra/smes/Report_18%20SMES.pdf<br />

(Accessed 08/11)<br />

Materials for Energy Storage Systems 38 © <strong>Granta</strong> <strong>Design</strong>, Jan 2012

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!