You are on page 1of 21

Compressed Air Energy Storage

Edward Barbour

In this unit you will be introduced to the Compressed Air Energy Storage concepts. These include
diabatic (conventional) CAES, Adiabatic CAES (ACAES) and Isothermal CAES (ICAES).

You will learn the general principle of each of these CAES variants and where they have been applied
or the state of development.

You will develop an understanding of the basic thermodynamic principles and be able to make some
estimates regarding system performance.

The notes are structured as follows:

1. Introduction & background


2. Thermodynamics of CAES systems
3. Different CAES systems and their analysis
4. Underground ACES
5. Challenges with compressed air energy storage systems for the future
6. Further Reading
7. References

Released along with these notes are a set of tutorial questions designed to build student
understanding. It is STRONGLY recommended that you work through the tutorial questions. Solutions
will be released the week after the energy storage module is taught.
1 Introduction and Background
Compressed air is broadly applied in industry where it is primarily used as an energy carrier for
applications which must operate in a spark-free environment. It is also used for cleaning, particularly
in the microelectronic sector. Within the industrial sector, it is generated almost exclusively by using
electrical energy to run compressors. However, in terms of energy supply outside of these special
industrial cases, it has never been established as an energy carrier. While early compressed air energy
carrying systems did exist for transferring power in some western European cities, these had many
drawbacks compared to electrical systems. In comparison to electricity, compressed air as an energy
carrier suffers from lower power and energy density and much higher losses.

Compressed air for energy storage, rather than as an energy carrier is a slightly different idea. The
general concept is that potential energy can be stored in the form of compressed air and released
when it is expanded. The first patents for compressed air energy storage systems date back to the
early 1940s.

However, while the ideas existed, actual progress in CAES remained limited until the 1960s. The
change was brought about by the large-scale deployments of nuclear power plants and large coal
plants as baseload generators. Suddenly, there was an economic case to store inexpensive off-peak
electric energy and use it at peak times. This period also coincided with large-scale developments of
Pumped Hydroelectric Energy Storage (PHES), however due to PHES’s strict geographical limitations,
CAES was also pursued as a suitable alternative for large-scale energy storage. The first CAES plant
was commissioned in 1969 in Germany, at a site called Huntorf [1]. Following this, and after a period
of heightened interest in the USA, the second CAES plant was built in McIntosh Alabama [2].

Despite the fact that both of these two plants are still in operation today, there has been no other
major plant development. This is largely down to two factors:

• Firstly, due to progress in Combined Cycle Gas Turbine (CCGT) technology that lead to
inexpensive gas power plants with high flexibility starting in the 1980s [3], the economic case
for large-scale high capital cost energy storage systems began to decline.
• Secondly, the existing CAES plants demonstrated are both hybrid air-storage/gas-combustion
plants, and so still have the disadvantages of emitting carbon dioxide and being reliant on the
price of gas for generation.

Fast forward to the renewable boom of the 2000’s (see Fig. 1) and there has again been heightened
interest in CAES technologies, as a potential large-scale storage option for low-cost intermittent
renewable generation. The majority of this work has focussed on CAES concepts which do not require
supplemental gas (or other fossil fuel) combustion. However, while there have been a number of R&D
projects undertaken since the mid-2000s, none have yet made it past the prototype stage.

There are a number of reasons why the performance of CAES has yet to reach its expected potential.
One reason may be that much of the development has occurred in small private ventures, where the
required times for return-on-investment are likely to be incompatible with the development of
medium-to-large scale mechanical systems. Furthermore, a high-profile European project with utility
partners, ADELE, also abandoned its developmental plans due to an uncertain economic environment
[4]. It must also be noted that at this time none of the liberalised electricity markets have seen the
development of any large-scale energy storage systems (there has been interest in large-scale energy
storage for power purposes – i.e. batteries for smoothing fluctuations in wind power – but these
systems have limited energy storage capacity). This includes Pumped Hydro, despite several plans for
new PHES systems being proposed over the last decade [3].

Figure 1: illustrating the timeline of CAES R&D and industrial efforts [5].

2 Thermodynamics of CAES systems


The basic concept of CAES can be understood using fundamental thermodynamics and in these notes,
we give an introduction to some of the fundamental thermodynamics of CAES systems. Our treatment
is necessarily limited given the practicalities of the course and interested students are encouraged to
consult the further reading.

Let us consider a simple example of a compressed air energy storage system. The system comprises a
compressor, an airtight container (the store) and a turbine as shown in Figure 2. To charge (to add
energy to) the system, electrical energy is used to drive the compressor (via a motor) which pushes
air into the container at High Pressure (HP). Once the system is pressurised to a sufficient degree, the
compressor is switched off and the air at high-pressure is contained within the HP store. When the
stored energy is required, the air is released from the store, driving a turbine in the process which can
be used to regenerate electricity via a generator.

Figure 2 : A simple CAES system


𝐸𝑙 𝐸𝑙
If the system requires electricity 𝑊𝑐ℎ to charge and returns electricity 𝑊𝑑𝑖𝑠 from the air expansion,
then the energy efficiency is appropriately:
𝐸𝑙
𝑊𝑑𝑖𝑠
𝜂= 𝐸𝑙 (1)
𝑊𝑐ℎ

To analyse this system, it is therefore clear that we must understand the work flows associated with
air compression and expansion. We will primarily be concerned with the mechanical-to-mechanical
𝐸𝑙
𝑊𝑑𝑖𝑠
efficiency, neglecting the efficiency of the motor and the generator (essentially assuming 𝜂 = 𝐸𝑙 ≈
𝑊𝑐ℎ
𝑊𝑑𝑖𝑠
).
𝑊𝑐ℎ

The work flows associated with a compression or expansion are best understood by considering a
control volume enclosing a compressor (or turbine). For students unfamiliar with a control volume,
further information can be found in most engineering thermodynamics textbooks, i.e. [6,7]. An
example of the control volume enclosing the compressor is shown in Figure 3.

Figure 3: A control volume enclosing a compressor

Using the conservation of energy, the energy balance for the control volume can be described by:

𝑑𝐸𝑐𝑣
= 𝑊̇𝑐𝑣 − 𝑄̇𝑐𝑣 + 𝑚̇1 (𝑢1 + 𝑝1 𝑣1 + 𝐾𝐸1 + 𝑃𝐸1 ) − 𝑚̇2 (𝑢2 + 𝑝2 𝑣2 + 𝐾𝐸2 + 𝑃𝐸2 ) (2)
𝑑𝑡

The terms in this equation are defined as follows: 𝐸𝑐𝑣 is the total energy contained within the control
𝑑𝐸𝑐𝑣
volume and therefore is the rate of total energy accumulation or reduction within the control
𝑑𝑡
volume. 𝑊̇𝑐𝑣 is the rate of useful work flowing into the control volume, 𝑄̇𝑐𝑣 is the heat output, 𝑢 is the
specific internal energy, 𝑝 is the pressure, 𝑣 is the specific volume and the numbers 1 and 2 denote
the inlet and exit respectively. The rate of mass flow into the control volume is given by 𝑚̇1 and the
mass flow leaving is 𝑚̇2 . (𝐾𝐸1 , 𝑃𝐸1 ) and (𝐾𝐸2 , 𝑃𝐸2 ) are the kinetic and potential energy at the inlet
and outlet states respectively.

In general, we assume that for a compressor (or expander) operating at steady state there is no build-
up of either mass or energy within the control volume. Therefore, the rate of total energy
𝑑𝐸𝑐𝑣
accumulation = 0 and since there is no rate of mass build up we deduce that 𝑚̇1 = 𝑚̇2 = 𝑚̇.
𝑑𝑡
Furthermore, in general the differences in the KE and PE terms are small compared to the differences
in the 𝑢 + 𝑝𝑣 terms, so Equation 2 reduces to Equation 3. The KE terms can generally be neglected for
compressors and turbines in stationary applications, however for propulsion devices – i.e. nozzles – we
cannot neglect the KE at the outlet.

𝑊̇𝑐𝑣 𝑄̇𝑐𝑣
− = ℎ2 − ℎ1 (3)
𝑚̇ 𝑚̇

Equation 3 is generally valid for compressors and expanders and should be remembered as a simplified
version of the conservation of energy for a compressor/expander control volume. It will be the starting
point for many questions on CAES systems.

In Equation 3, we have grouped the 𝑢 + 𝑝𝑣 terms together into the combination property enthalpy,
ℎ.

ℎ = 𝑢 + 𝑝𝑣 (4)

Enthalpy is not directly measurable. The best way to calculate enthalpy ℎ is to use tabulated values
that are given in many thermodynamic textbooks or reference softwares depending on the pressure
and temperature of the gas. This in turn will allow us to calculate the work or heat input required if
we know the other terms. So, for example, if we measured the heat production during a compression
and the temperature and pressure at the inlet and exit states (all quantities that we could practically
measure), we could look up the enthalpy values at the particular pressures and temperatures to
estimate the compressor work per unit mass flow.

The situation can be further simplified if 𝑄̇𝑐𝑣 = 0, as then the compressor work will just depend on
the enthalpy at the inlet and exit. In general, there will always be some heat flow associated with any
temperature difference. However, if the compression happens sufficiently quickly then the heat
escaping per unit mass flow may be negligible. This way, equation (3) can be further simplified to

𝑤̇𝑐𝑣
= ℎ2 − ℎ1 (5)
𝑚̇

The aforementioned thermodynamic concepts presented are used to analyse the charging stage of
simple Compressed Air Energy Storage systems. By adapting equations (1) to (5) – rearranging the
signals of work and heat flows in accordance to thermodynamic convention – they are also valid for
analysing the discharge process. Specific aspects will be now discussed with greater detail.

2.1 Ideal gas law

Up until this point, we have made statements that are generally true for most gases. However, when
we make the ideal gas assumption, the situation can be simplified further.

The ideal gas law makes several key assumptions:

• The gas molecules move randomly and collisions with the walls of the container and between
the molecules are perfectly elastic
• The gas molecules occupy a much smaller volume than the total volume occupied by the gas
• The particles do not interact with attractive or repulsive intermolecular forces.
• The average kinetic energy of the gas particles is directly proportional to the temperature.

Although this may seem quite restrictive, it is actually representative of a wide range of conditions
typically encountered. It breaks down when the temperature is very low and at very high pressures,
since in these extreme environments intermolecular forces and the volume occupied by the gas
molecules become important. Real gas models relax these two assumptions by providing a correction
for the volume occupied by the molecules and the intermolecular forces.

The ideal gas law states:

𝑝𝑉 = 𝑛𝑅∗ 𝑇 (6)

This should be committed to memory and is the starting point for many analysis problems.

The terms in Equation 6 are as follows: 𝑝 is the gas pressure in Pascals (Pa), 𝑉 is the volume it occupies
in m3, 𝑛 is the amount of gas in moles and 𝑇 is the temperature in Kelvin (K). The term 𝑅∗ is the
Universal gas constant, 𝑅∗ = 8.314 kJmol−1 . Equation 6 states the molar form of the ideal gas law. In
general, for engineering applications, it is more useful to work with mass than amounts in moles.
Therefore, we can re-express the ideal gas law and use, 𝑅, the specific gas constant with units of Jkg −1 .

𝑝𝑉 = 𝑚𝑅𝑇 (7)

𝑅 is specific to the exact gas in question, and can be calculated by dividing the Universal gas constant
by the Molar Mass of the gas. For example, the molar mass of air 𝑀𝑀𝑎𝑖𝑟 = 0.02897 kgmol−1 (in this
case air is consider a mixture composing of 78% 𝑁2 ; 21% 𝑂2 & 1% 𝐴𝑟 𝑀/𝑀). Dividing the Universal
gas constant 𝑅∗ by 𝑀𝑀𝑎𝑖𝑟 yields the specific gas constant for air, 𝑅𝑎𝑖𝑟 = 287 Jkg −1 .

For an ideal gas, pressure and volume can be related to each other during a compression or expansion
by the Equation 8.

𝑝𝑉 𝑛 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 (8)

There are two specific cases of interest.


𝑐𝑝
1. 𝑛 = 𝛾 = . In this case the compression is adiabatic and we have that 𝑄̇𝑐𝑣 = 0 and 𝑝𝑉 𝛾 =
𝑐𝑣
𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡. 𝛾 (sometimes also presented as 𝑘) is the heat capacity ratio or specific heat ratio.
This can be proved by considering the first law of thermodynamics and the ideal gas law
together and considering an infinitesimal adiabatic process for which 𝑑𝑄 = 0. The steps are
to differentiate the ideal gas law and note that 𝑐𝑝 − 𝑐𝑣 = 𝑅. This is left as an exercise and is
beyond the scope of the discussion here.
2. 𝑛 = 1. In this case the compression/expansion is isothermal, because the ideal gas law states
𝑝𝑉 = 𝑚𝑅𝑇 (and R is just a constant specific to each gas).
2.2 Adiabatic compression and expansion

For the case of adiabatic compression or expansion, we have that the compression must happen
quickly enough that we can neglect the heat transfer out of the gas to the surroundings. In practice,
this a good approximation for many compressions and expansions that happen in industrial processes
and for energy and power applications (for example industrial compressions and gas turbines for
power generation). It is also most applicable for compressors in energy storage systems.

Since for the adiabatic case, 𝑝𝑉 𝛾 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡, we can also write:

𝑝𝑉 𝛾 = 𝑝1 𝑉1𝛾 = 𝑝2 𝑉2𝛾 (9)

Then, using the ideal gas law, we can eliminate the volume term to express this in terms of pressure
and temperature as the only variables:

𝑚𝑅𝑇1 𝛾 𝑚𝑅𝑇2 𝛾
𝑝1 ( ) = 𝑝2 ( ) (10)
𝑝1 𝑝2

Therefore, 𝑝11−𝛾 𝑇1𝛾 = 𝑝21−𝛾 𝑇2𝛾 and:


𝛾−1
𝑝 𝛾
𝑇2 = (𝑝2 ) 𝑇1 (11)
1

Note that for a compression, i.e. where 𝑝1 < 𝑝2, the temperature 𝑇1 < 𝑇2 , whereas for an expansion,
where 𝑝1 > 𝑝2, we find the temperature after expansion is lower, i.e. 𝑇1 > 𝑇2 .Values for 𝑐𝑝 and 𝑐𝑣
(and consequently 𝛾) can be found in thermodynamic tables. Albeit being dependent on temperature
conditions, 𝛾 = 1.4 is a reasonable approximation over a wide range of conditions (200 < 𝑇 <
1000𝐾 returns 1.401 < 𝛾 < 1.336).

2.3 Isentropic compressor and turbine efficiency

Adiabatic compressions and expansions are reversible and are examples of isentropic processes,
meaning that these processes happen at constant entropy. Hence, the minimum work required to
compress air from 𝑝1 to 𝑝2 is denoted as 𝑊𝑖𝑠 .

A full discussion of entropy, denoted 𝑠, is outside the scope of the notes here, but it is given extended
treatments in almost every engineering thermodynamics textbook, i.e. [6,7]. It is a function of state of
the system, that is it depends only on the current state of the system, not how it came to be and is a
consequence of the second law of thermodynamics. In engineering terms, it is useful as when a system
undergoes a change in energy, there is some transfer of unusable heat to the surroundings 𝑄 = 𝑇𝑎 Δ𝑆
which cannot be converted to work. In addition, we find that for real processes there is a quantity of
lost work given by 𝑊𝑙𝑜𝑠𝑡 = 𝑇𝑎 𝑆𝑖 , where 𝑆𝑖 is the entropy production due to irreversible processes. The
second law demands that 𝑆𝑖 ≥ 0. Isentropic processes are those for which 𝑆𝑖 = 0.

The operation of real machinery is not isentropic, but typically introduces some extra heat into the
compression or expansion from friction in the machinery. Therefore, if there is no heat loss to the
surroundings, the temperature rise in the air may be greater than for the purely isentropic case. We
define the isentropic compressor efficiency as:
𝑐𝑜𝑚𝑝 𝑊̇𝑖𝑠
𝜂𝑖𝑠 = 𝑊̇ (12)
𝑎𝑐𝑡𝑢𝑎𝑙

In Equation 12 𝑊̇𝑖𝑠 is the work rate in the isentropic case and 𝑊̇𝑎𝑐 is the actual work rate given the
irreversibility present.

Similarly, for a turbine, the friction in the machine increases the outlet temperature and thus
decreases the work available from expanding the air. The isentropic turbine efficiency is defined as:

𝑡𝑢𝑟𝑏 𝑊̇𝑎𝑐𝑡𝑢𝑎𝑙
𝜂𝑖𝑠 = (13)
𝑊̇𝑖𝑠

For the specific case of air at the temperatures and pressures we generally encounter, we can assume
that it behaves like an ideal gas, and therefore the enthalpy ℎ is only a function of temperature (this
is shown by noting that the internal energy 𝑢 is a function of temperature only and combining this
with the ideal gas equation and the definition for enthalpy). Therefore, the enthalpy ℎ is just:

ℎ = 𝑐𝑝 𝑇 (14)

Furthermore, if the range of temperatures encountered is moderate, we can make the perfect gas
approximation that assumes a constant specific heat capacity 𝑐𝑝 .

Then, we can directly calculate the work by substituting Equation 14 into Equation 5 and using
Equation 11. This yields:
𝛾−1
𝑊̇𝑐𝑣 𝑝 𝛾
= 𝑐𝑝 𝑇1 ((𝑝2 ) − 1) (15)
𝑚̇ 1

Note that the work rate is positive when 𝑝2 is greater than 𝑝1 and negative when the other way
around. Hence the system requires work added when it is being compressed and does work on the
surroundings when being expanded.

Since with constant specific heat capacity, the work depends only on temperature, we can then
combine Equations 15 and 12 to express the isentropic compressor efficiency in terms of temperatures
only:

𝑐𝑜𝑚𝑝 ℎ2,𝑖𝑠 −ℎ1 𝑇2,𝑖𝑠 −𝑇1


𝜂𝑖𝑠 = =𝑇 (16)
ℎ2 −ℎ1 2,𝑎𝑐𝑡𝑢𝑎𝑙 −𝑇1

𝛾−1
𝑝 𝛾
Where 𝑇2,𝑖𝑠 is given by Equation 11, i.e. 𝑇2,𝑖𝑠 = (𝑝2 ) 𝑇1 . We can then rearrange this to get the actual
1
temperature 𝑇2,𝑎𝑐𝑡𝑢𝑎𝑙 :
𝛾−1
𝑝
𝑇1 ((𝑝2) 𝛾 −1)
1
𝑇2,𝑎𝑐𝑡𝑢𝑎𝑙 = 𝑐𝑜𝑚𝑝 + 𝑇1 (17)
𝜂𝑖𝑠

For the expander or turbine, the temperatures are related to the isentropic efficiency by:

𝑡𝑢𝑟𝑏 𝑇1 −𝑇2,𝑎𝑐𝑡𝑢𝑎𝑙
𝜂𝑖𝑠 = (18)
𝑇1 −𝑇2,𝑖𝑠
Therefore, the actual temperature at the turbine outlet is:
𝛾−1
𝑡𝑢𝑟𝑏 𝑝 𝛾
𝑇2,𝑎𝑐𝑡𝑢𝑎𝑙 = 𝑇1 − 𝜂𝑖𝑠 𝑇1 (1 − (𝑝2 ) ) (19)
1

2.4 Heat Exchangers

As we have seen in the previous Equations, the exhaust air from compressors is hotter than the inlet
air while the reverse is true for expanders (like turbines). There are a number of reasons that mean
that we do not want to work with hot air. In particular, hot air is volumetrically inefficient to store and
mechanically difficult to work with.

Therefore, in CAES systems the air is typically cooled between the compressions. Each compressor
also normally has the same (or similar) compression ratios since the goal of staging the compression
is to minimise the temperature of the air at the compressor outlets.

In this course we consider counter-current heat exchangers, with a simple diagram of a counter
current heat exchanger shown in the lower diagram in Figure 4.

Figure 4: Concurrent and counter-current heat exchangers.

For heat exchangers we define an effectiveness, which is the ratio of the heat transfer rate achieved
versus the theoretical maximum, as shown in Equation 19.

𝑄̇
𝜖 = 𝑄̇ (20)
max

The maximum heat transfer possible would involve raising the cold fluid to the inlet temperature of
the hot fluid or cooling the hot fluid to the inlet temperature of cold fluid. Assuming constant specific
heats for both the hot and cold fluid, the change in the temperature of the hot fluid is given by Δ𝑇ℎ =
𝑄̇
, where 𝑚̇ is the fluid mass flow rate and 𝑐 is the heat capacity. Similarly, the temperature change
𝑚̇ ℎ 𝑐ℎ
𝑄̇
in the cold fluid is Δ𝑇𝑐 = 𝑚̇ .
𝑐 𝑐𝑐

For a counter-current heat exchanger where the fluid heat flow rates (the mass flow rate times the
heat capacity) are matched, i.e. 𝑚̇ℎ 𝑐ℎ = 𝑚̇𝑐 𝑐𝑐 , the maximum heat transfer 𝑄̇𝑚𝑎𝑥 = 𝑚̇𝑐 𝑐𝑐 (𝑇𝑐,𝑜𝑢𝑡𝑙𝑒𝑡 −
𝑇𝑐,𝑖𝑛𝑙𝑒𝑡 ) = 𝑚̇ℎ 𝑐ℎ (𝑇ℎ,𝑖𝑛𝑙𝑒𝑡 − 𝑇ℎ,𝑜𝑢𝑡𝑙𝑒𝑡 ). Therefore, we can write the effectiveness 𝜖 as expressed in
Equation 19.
𝑇ℎ,𝑖𝑛−𝑇ℎ,𝑜𝑢𝑡 𝑇𝑐,𝑜𝑢𝑡 −𝑇𝑐,𝑖𝑛
𝜖= = (21)
𝑇ℎ,𝑖𝑛−𝑇𝑐,𝑖𝑛 𝑇ℎ,𝑖𝑛 −𝑇𝑐,𝑖𝑛

Using this, if we are given the inlet fluid temperatures and the heat exchanger effectiveness, we can
calculate the outlet fluid temperatures.

In reality, there will also always be some pressure decrease through a heat exchanger which we denote
as Δ𝑝𝑑𝑟𝑜𝑝 .

2.5 Isothermal Compression

For the 𝑛 = 1 case, we have that 𝑝𝑉 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡. Reconciling this with the ideal gas law 𝑝𝑉 = 𝑚𝑅𝑇
we can see that 𝑇 must be constant in this case, since 𝑚 is the mass (which is by definition constant
for a fixed mass of gas) and 𝑅 is a constant. For an ideal gas, ℎ is only a function of temperature, and
Equation 2 then shows us that 𝑊̇𝑐𝑣 = 𝑄̇𝑐𝑣 . Therefore, in an isothermal compression, the work that is
put into the compression manifests itself as heat which must be reversibly lost to the surroundings.

We can then calculate the work stored when a volume of air is compressed isothermally. First, since
we have that ℎ1 = ℎ2 in Equation 2 and we know that the heat term is non-zero, we use the definition
of work:

d𝑊 = −𝑝𝑑𝑉 (22)

Here, a positive 𝑊 means that work is done on the system and negative 𝑊 means the system does
𝑚𝑅𝑇
work on the surroundings. Using the ideal gas law, we can express the pressure 𝑝 = . Substituting
𝑉
𝑑𝑉
this into the work equation yields 𝑑𝑊 = −𝑚𝑅𝑇 ( 𝑉 ). Now integrating this from state 1 to state 2
gives:

𝑉 𝑑𝑉 𝑉
∫ 𝑑𝑊 = −𝑚𝑅𝑇 ∫𝑉 2 = −𝑚𝑅𝑇 ln (𝑉2 ) (23)
1 𝑉 1

Therefore, we see that if 𝑉2 is smaller than 𝑉1 (i.e. the pressure is higher), then work is required. We
can rearrange this using the ideal gas law to give:
𝑚𝑅𝑇
𝑝2 𝑝 𝑝 𝑝
𝑊 = −𝑚𝑅𝑇 ln ( 𝑚𝑅𝑇 ) = 𝑚𝑅𝑇 ln ( 2 ) = 𝑚𝑅𝑇 ln ( 2 ) = 𝑝1 𝑉1 ln ( 2 ) (24)
𝑝1 𝑝1 𝑝1
𝑝1
3 Different Compressed Air Energy Storage systems and their analysis

Figure 5: characterisation of CAES systems

3.1 Diabatic CAES systems

Existing CAES plants are also known as conventional (also known as diabatic) plants. These are
essentially gas turbines in which air is pre-compressed using off-peak electricity, rather than running a
turbine and compressor simultaneously. In these plants, off-peak cheaper grid electricity is used to
compress air which is stored, and then mixed with natural gas and combusted during expansion.
Compression is staged and the majority of the compression heat wasted. Currently there are two
commercial CAES plants worldwide; the Huntorf plant in Germany and the McIntosh plant in Alabama.

• Huntorf CAES plant: Data from [10]. 310,000m3 cavern at a depth of 600m, pressure tolerance
between 50 - 70 bar, converted from a solution mined salt dome. Daily charging cycle of 8h,
output of 290MW for 2 hours. 0.8kWh of electricity and 1.6kWh of gas required to produce
1kWh of electricity. Notably, built when the price of gas turbines was historically high.
• McIntosh CAES plant: Data from [11]. 538,000m3 salt cavern at a depth of 450m, pressure
tolerance between 45-76 bar. Originally it provided an output of 110MW for 26 hours but in
1998 two extra generators were added and its total output capacity is now 226MW. 0.69kWh
of electricity and 1.17kWh of gas to produce 1kWh of electricity.

Both plants are commercially viable and still running in their respective markets!
Figure 6: Schematic of diabatic CAES system.

Figure 6 shows a simplified diagram of a diabatic CAES system. In this system, we can calculate the
work rate required to charge the system based on the enthalpy difference across the compressors.
Therefore, the work rate required is

𝑊̇𝑐ℎ = 𝑚̇𝑐ℎ (ℎ2 − ℎ1 + ℎ4 − ℎ3 + ℎ6 − ℎ5 ) (25)


𝑤𝐶1 𝑤𝐶2 𝑤𝐶3

To calculate the total energy required for charging the system, we must know the values of these
quantities at all times. In a store with a fixed volume, its pressure will increase as the system is charged.
Therefore, the enthalpies and the mass flow rate may also change. This means we can only easily
calculate the work if the mass flow rate and enthalpies are constant throughout the charge time.

If the mass flow rate and enthalpies are constant then:

𝑊𝑐ℎ = 𝑊̇𝑐ℎ 𝜏𝑐ℎ (26)

Where 𝜏𝑐ℎ is the charge duration.

if we make the perfect gas assumption then the specific enthalpies at the compressor/expander
outlets can be calculated from the inlet pressure, compressor/expander pressure ratio (𝑟), inlet
temperature and compressor/expander isentropic efficiency. For example, under the perfect gas
assumption the enthalpy ℎ2 would be given by:

𝛾−1
𝛾
((𝑟1 )−1)

ℎ2 = 𝑐𝑝 𝑇1 1 + (27)
𝜂𝑐

[ ]

where 𝑟1 is the pressure ratio of the first compressor. The temperature at the inlet of compressor 2
must be calculated considering the effectiveness of the intercooling heat exchanger 𝐼𝐶1 and the
temperature of the cool fluid inlet – in this case the ambient temperature.

𝑇3 = 𝑇2 − 𝜖𝐼𝐶1 (𝑇2 − 𝑇𝑎𝑚 ) (28)

The work out from expansion is calculated by considering the enthalpy difference across the turbines,
so that for the system shown in Figure 6 the work rate out is:

𝑊̇𝑒𝑥 = 𝑚̇𝑒𝑥 (ℎ9 − ℎ10 + ℎ11 − ℎ12 ) (29)


𝑤 𝑇1 𝑤 𝑇2
Again, this is a work rate and can only be used to calculate the total discharging work if the mass flow
rate and enthalpies are constant.

A key difference between diabatic and adiabatic CAES is the presence of combustion chambers or pre-
heaters with an external heat source. In the current analysis, we treat these as heaters supplying heat
rates 𝑄̇1 and 𝑄̇2 to the air flowing through. Using the conservation of energy, Equation 3, we deduce
that the change in enthalpies ℎ8 and ℎ9 is due to the heat rate 𝑄1̇ , since there is no work extracted
from the combustion-chamber/heater. Therefore:

𝑄̇1 = 𝑚̇𝑒𝑥 (ℎ9 − ℎ8 ) (30)

Similarly, we deduce that 𝑄̇2 = 𝑚̇𝑒𝑥 (ℎ11 − ℎ10 ).

Using the above approach is it possible to work out the rates of work input during charge and work
output during discharge. It is also possible to calculate the total work in and work out if the storage
pressure (and in turn the mass flow rates and enthalpies) are constant.

Remarks

It is interesting to note that despite the strong resemblance of diabatic CAES to gas turbines, the inlet
pressures for the HP turbine are higher than for typical gas turbines. In Huntorf, the inlet pressure is
44 bar which is much higher than the equivalent for a typical gas turbine (about 11 bar). The result is
the HP turbine at Huntorf is based on a small-intermediate steam turbine design.

The throttle valve (𝑅) or regulator shown in the Figure 6 is used to regulate the pressure entering the
HP turbine, ensuring that the system can operate at steady-state (which in turn increases the
performance and simplicity of the turbine design). While throttle valves are constant enthalpy devices
– they do introduce irreversibility. This is due to the entropy change that occurs as the gas expands
through the device.

This is often confusing, since the change in enthalpy gives the work – but the crucial thing to note is
that while the enthalpy at the outlet of the throttle valve doesn’t change – the enthalpy that it can
reach after expansion has changed! Hence, a throttle valve is always a source of irreversibility that
represents lost work in an ideal system.
3.2 Adiabatic Compressed Energy Storage (ACAES) systems

Figure 7: Schematic of an example ACAES configuration [9]. It should be noted that ACAES systems
may or may not include a throttle valve for maintaining a constant expansion pressure.

In adiabatic CAES systems (such as that shown in Figure 7), the heat that is removed from the air
between the compressions and after the last compression is stored and returned to the air during the
expansion process. It must be stored in separate Thermal Energy Stores (TESs). This produces an
acceptable storage efficiency without the need for supplemental combustion.

Since the compressor outlet temperatures may be different for the different compression stages,
there may be a separate thermal storage for each compression stage to avoid mixing heat at different
temperatures. For example, if the heat exchanger effectiveness 𝜖<1, then the temperature of the air
entering the second compressor C2 will be higher than the air entering the first compressor C1 .
Therefore, the outlet temperatures will also be different, as will the temperatures of the heat
exchange fluids entering the thermal stores TES1 and TES2 . From our knowledge of thermal energy
storage, we know that mixing of heat stored at different temperatures is a source of irreversibility and
lost work.

The analysis of these systems is very similar to that which was described previously for CAES systems.
Therefore, we can deduce that the work rate required for the charging process is:

𝑊̇𝑐ℎ = 𝑚̇𝑐ℎ (ℎ2 − ℎ1 + ℎ4 − ℎ3 + ℎ6 − ℎ5 ) (31)

Similarly, the work rate available from the discharging process is:

𝑊̇𝑒𝑥 = 𝑚̇𝑒𝑥 (ℎ9 − ℎ10 + ℎ11 − ℎ12 + ℎ13 − ℎ14 ) (32)


As with diabatic CAES systems, if we make the perfect gas assumption then the specific enthalpies at
the compressor/expander outlets can be calculated from the inlet pressure, compressor/expander
pressure ratio, inlet temperature and compressor/expander isentropic efficiency.

The temperature entering each the thermal stores is given by the outlet temperature of the cold fluid
stream in each of the heat exchangers. For counter-current exchangers with equal heat flow rates,
𝑚̇ℎ 𝑐ℎ = 𝑚̇𝑐𝑐 , this can be calculated by considering the cold fluid streams in the heat exchangers and
the exchanger effectiveness. For example, the temperature entering the TES1 is given by:

𝑇𝑇𝐸𝑆1 = 𝑇𝑐𝑖 + 𝜖(𝑇2 − 𝑇𝑐𝑖 ) (33)

where 𝑇𝑐𝑖 is the cold fluid inlet temperature, which in this case is equal to the ambient 𝑇𝑎 .

As with CAES systems, unless there is some way of maintaining the pressure in the store at a constant
level throughout the charging/discharging process, the pressure will increase during charging and
decrease during discharging. Thus, the temperature entering the TESs will constantly increase
throughout the compression process. The final temperature of the TESs will then depend on the
degree of mixing in the TES, with a perfectly mixed TES reaching equilibrium at the average
temperature minus any losses. It is beyond the scope of the discussion here to do a full consideration
of the thermodynamics of the TES, but during the analysis of the expansion process care must always
be taken to ensure that any specified temperature drops are accounted for.

Remarks

It is apparent that the expansion in ACAES systems can only produce the same work as the
compression if both the compression and expansion follow reversible paths. The simplest way to do
this is if the expansion process follows the exact reverse of the compression process. Therefore, for
ACAES systems, designs in which the expansion process is the exact reverse of the compression
process can have a theoretical efficiency limit of 100%. For example, if you have 3 compression stages
with cooling after each stage, to achieve a theoretical efficiency limit of 100% there should also be 3
expansion stages with heating prior to each expansion.

While ACAES is a well-established concept, there has been no successful demonstration of the
technology to date. Recent prototype systems in Switzerland [20] and China [21] are aiming to
challenge this. Some of the major development challenges for ACAES systems are discussed in Section
5.

3.3 Isothermal CAES systems

Isothermal CAES systems aim to reduce the need for high temperature TES, thus potentially making
the system more cost-effective. They could also potentially achieve very high compression ratios,
negating the need for multiple compressions and expansions. Isothermal compressions and
expansions are also particularly attractive as they are highly reversible, since heat is transferred across
very small temperature differences and hence the lost work is very small. A simple isothermal CAES
concept is shown below.
Figure 8: A simple isothermal CAES system [8]

For the system in Figure 8 we can calculate the work stored as the pressure in the store changes from
𝑀𝑋 . In this case, we know that the work
the ambient pressure 𝑝0 to some maximum storage pressure 𝑝𝑠𝑡
required to compress each increment of air isothermally (which is then added to the store) is given by
Equation 24, expressed for a mass increment as shown in Equation 34:

𝑝
𝛿𝑊 = 𝛿𝑚𝑠𝑡 𝑅𝑇0 𝑙𝑛 ( 𝑝𝑠𝑡 ) (34)
0

𝑀𝑋 ) and the temperature is constant


Here 𝑝𝑠𝑡 is the pressure in the store (which it changes from 𝑝0 to 𝑝𝑠𝑡
at 𝑇0 (the store is also considered isothermal). Since the compression is isothermal, we also know how
the pressure in the store changes as more mass is added at constant temperature, as the gas in the
storage volume 𝑉𝑠 is described by the ideal gas law. Therefore, in the HP air store we have 𝑝𝑠 𝑉𝑠 =
𝑉
𝑚𝑠𝑡 𝑅𝑇0 and 𝛿𝑚𝑠 = 𝑅𝑇𝑠 𝛿𝑝𝑠𝑡 .
0

Substituting 𝛿𝑚𝑠𝑡 into the Equation 34 allows us to perform the integration:

𝑝𝑀𝑋 𝑝𝑀𝑋
∫ 𝑑𝑊 = 𝑉𝑠 ∫𝑝 𝑠𝑡 ln 𝑝𝑠𝑡 − ln 𝑝0 𝑑𝑝𝑠𝑡 = 𝑉𝑠 [𝑝𝑠𝑡 ln 𝑝𝑠𝑡 − 𝑝𝑠𝑡 − 𝑝𝑠𝑡 ln 𝑝0 ]𝑝0𝑠𝑡 (35)
0

Putting in the limits yields:


𝑀𝑋
𝑀𝑋 𝑉 (ln ( 𝑝𝑠𝑡 𝑝0
𝑊 = 𝑝𝑠𝑡 𝑠 )− 1 + 𝑝𝑀𝑋 ) (36)
𝑝0 𝑠𝑡

If the system is 100% efficient, then Equation 36 also gives the available work which can be extracted
from the system.

Remarks

In isothermal CAES the compressions aim to be as close to isothermal as possible. This is achieved by
minimising the temperature differences which drive heat flow from the compressors to the
environment (which is at a lower temperature). A huge challenge here is to make an isothermal
compression process which operates sufficiently quickly to be of practical industrial importance but
which maintains the small temperature differences required for high reversibility. One idea for near-
isothermal compression which has been suggested by LightSail (a start-up company in California)
involves a water spray into the compression chamber of a specially designed reciprocating
compressor/expander unit (see Figure 9). The water droplets absorb the heat of compression and
their high specific heat capacity causes the temperature increase in the compression chamber to be
much smaller. This warm water is then stored and on discharge is re-injected as a mist into the
reciprocating machine which now acts as an expander. Liquid piston designs have also been proposed.

Figure 9: Illustrating a near-isothermal CAES concept [15]

Near-Isothermal CAES was also being pioneered by SustainX who tried to build a 1.5MW prototype,
however this company has ceased operations citing spiralling system costs. Lightsail Energy has also
shut down operations as of 2018.

4 Underground CAES
As with Pumped Hydro Storage (PHS), CAES which uses underground air storage also requires
favourable geography to provide the underground air storage caverns. However, there are many more
suitable sites worldwide than for PHS, although the costs are highly site specific. For example, in the
UK, a recent (2019) paper estimated that there was sufficient suitable CAES sites in the UK to store
160% of the UK’s winter electricity demand [16].

In terms of costs, mining a suitable underground cavern where suitable geology doesn't exist or
creating an above-ground equivalent storage container may be prohibitive depending on the exact
geology. However, a naturally occurring cavern or somewhere easily minable may offer a very
attractive price of storage in terms of $/kWh (or dollars per metre cubed of air storage).

Caverns can be created in salt geology (typically using salt solution mining techniques) or existing
caverns can be exploited provided that they are capable of housing the desired pressure. Geological
formations such as aquifers and salt formations (bedded salt and domal salt) offer potential locations.
Costs can also be reduced if existing well infrastructure is in place from previous underground drilling
operations- for example from the petrochemical industry. While specific geology is required, this
geology is relatively widespread. For example, the EPRI suggests that up to 80% of the US could have
favourable geology [17] (see Figure 10).

Figure 10: US geology for compressed air caverns. Regions with high wind resources are also indicated
with the idea that CAES sites and wind turbines could be co-located [18].

Estimates for the costs of cavern mining can be as low as $1/kWh of storage capacity if solution mining
techniques can be used [19]. In solution mining, fresh water is pumped in a salt deposit, becomes
saturated with salt and is then removed as brine. One further problem however is that disposal of this
brine can cause environmental issues.

5 Challenges with compressed air energy storage systems for the future
There are several challenges which must be overcome before adiabatic CAES can become a viable
energy storage technology option.

• Specialised compressor equipment must be developed, in which the heat generated during
the compression procedure is stored in a highly reversible manner. This process currently
seems most likely to consist of a series of adiabatic compressions in which heat losses from
the compressor to the surroundings are minimised, although significant work is being done
on near-isothermal compressor/expander development. The compressors must also operate
with much higher compression ratios than current compressors which do not involve cooling
during the compression. Each of the compressions is then followed by a cooling stage which
aims to reversibly extract the compression heat. Possible options for heat extraction include
packed bed regenerators or counter-current indirect contact air-to-fluid heat exchangers.
This type of compression equipment is fundamentally different to many industrial
compressors. Why? Because the vast majority of compressors are designed to minimise the
work required to achieve air at a given pressure. Most industrial compressions then typically
involve trying to shed as much heat as possible from the compression process - as hot air
takes more work to compress. The ACAES process is fundamentally different as reversibility
should be maximised rather than work minimised.
• Specialised expansion equipment must also be developed. Air turbines which provide highly
isentropic expansions and operate within the desired pressure ratios are required. While
these turbines do not currently exist on the industrial market, it is anticipated that their design
can learn much from the current generation of gas turbines for power generation. The
pressure ratios will likely be smaller than most current gas turbines, although the absolute
inlet pressure in the HP expander will be higher. One specific advantage is that the material
demands will be much less (in terms of temperature tolerance) than current gas turbines
which operate with inlet temperatures up to 2200K. Certainly the development of efficient
reversible compression/expansion infrastructure which be a huge benefit, analogous to the
reversible pump/turbines that exist in Pumped Hydro.
• Sliding pressures. Unless the system can be operated between constant operational
pressures, both the compression and expansion machinery must operate at high efficiency
over a range of pressure ratios. Furthermore, it is very difficult to make effective isothermal
air storage containers, therefore a further sizeable loss in CAES systems relates to the
temperature increase in the HP air store as the pressure is increased. This leads to a reduced
air mass which can be added to the store, and during the expansion process significantly
restricts the amount of air that can be effectively removed from the store by causing the
pressure to drop as the air in the store cools. A single constant high-pressure air storage is a
primary advantage of UnderWater CAES.
• High pressure air storage. Depending on the chosen method of storage high pressure, air
storage tanks must be developed which have minimum cost and/or technology for lining
caverns for minimal leakage needs to be developed.
• Highly reversible heat exchangers are required which can minimise the temperature
difference between the working fluid and the thermal storage medium while introducing
minimal pressure drops.

6 Further reading
Students wishing for further information are strongly recommended to read the excellent review
articles by Budt et al. [5] and Kim et al. [8].

The following thermodynamics textbooks are recommended and provide a good introduction to
engineering thermodynamics at an undergraduate level.

• Moran and Shapiro, Fundamentals of Engineering Thermodynamics


• Cengel and Boles, Thermodynamics: An Engineering Approach

More advanced texts:

• Bejan, Advanced Engineering Thermodynamics


7 References
[1] Crotogino F, Mohmeyer K-U, Scharf R. Huntorf caes: more than 20 years of successful operation.
Spring 2001 Meeting, Orlando, Florida, USA; 15–18 April 2001.

[2] Nakhamkin M, Chiruvolu M. Available Compressed Air Energy Storage (CAES) Plant Concepts; 2007.

[3] Barbour, E., Wilson, I. G., Radcliffe, J., Ding, Y., & Li, Y. (2016). A review of pumped hydro energy
storage development in significant international electricity markets. Renewable and Sustainable
Energy Reviews, 61, 421-432.

[4] RWE Power. ADELE - Adiabatic Compressed Air Energy Storage for Electricity Supply.
https://www.rwe.com/web/cms/mediablob/en/391748/data/364260/1/rwe-power-
ag/innovations/Brochure-ADELE.pdf

[5] Budt, M., Wolf, D., Span, R., & Yan, J. (2016). A review on compressed air energy storage: Basic
principles, past milestones and recent developments. Applied Energy, 170, 250-268.

[6] Moran, M. J., Shapiro, H. N., Boettner, D. D., & Bailey, M. B. (2010). Fundamentals of engineering
thermodynamics. John Wiley & Sons.

[7] Cengel, Y. A., & Boles, M. A. (2002). Thermodynamics: an engineering approach. Sea, 1000, 8862.

[8] Kim, Y. M., Lee, J. H., Kim, S. J., & Favrat, D. (2012). Potential and evolution of compressed air
energy storage: energy and exergy analyses. Entropy, 14(8), 1501-1521.

[9] He, W., Luo, X., Evans, D., Busby, J., Garvey, S., Parkes, D., & Wang, J. (2017). Exergy storage of
compressed air in cavern and cavern volume estimation of the large-scale compressed air energy
storage system. Applied Energy, 208, 745-757.

[10] BBC Brown Boveri. Huntorf Air Storage Gas turbine Power Plant.
https://www.eon.com/content/dam/eon-content-pool/eon/company-asset-finder/asset-
profiles/shared-ekk/BBC_Huntorf_engl.pdf

[11] M. Nakhamkin, L. Andersson, E. Swensen, J. Howard, R. Meyer, R. Schainker, R. Pollak, and B.


Mehta, J. Eng. Gas Turbines Power 114, 695 (1992). https://doi.org/10.1115/1.2906644

[12] Kaldellis, J. K. & Zafirakis, D., 2007. Optimum energy storage techniques for the improvement of
renewable energy sources-based electricity generation economic efficiency.. Energy, Volume 32, p.
2295–2305.

[13] Chen, H. et al., 2009. Progress in electrical energy storage system: A critical review. Progress in
Natural Science, Volume 19, pp. 291-312.

[14] EPRI, 2010. Electricity Energy Storage Technology Options.


http://large.stanford.edu/courses/2012/ph240/doshay1/docs/EPRI.pdf

[15] Fong, D. Insights by Danielle Fong. https://daniellefong.com/


[16] Mouli-Castillo, J., Wilkinson, M., Mignard, D., McDermott, C., Haszeldine, R. S., & Shipton, Z. K.
(2019). Inter-seasonal compressed-air energy storage using saline aquifers. Nature Energy, 1.

[17] Compressed Air Energy Storage: Renewable Energy (2010, March 17) retrieved 22 April 2017 from
https://phys.org/news/2010-03-compressed-air-energy-storage-renewable.html

[18] Succar, S & Williams, R.H.. Compressed Air Energy Storage: Theory, Resources, and Applications
for Wind Power, Princeton University (published April 8, 2008)

[19] De Samaniego Steta, F. Modeling of an Advanced Adiabatic Compressed Air Energy Storage (AA-
CAES) Unit and an Optimal Model-based Operation Strategy for its Integration into Power Markets.
EEH – Power Systems Laboratory. Swiss Federal Institute of Technology (ETH) Zurich

[20] Synthesis of the NRP 70 joint project “Electricity storage via adiabatic air compression”
https://nfp-energie.ch/en/dossiers/191/

[21] Compressed Air Energy Storage: The Path to Innovation http://en.cnesa.org/latest-


news/2019/9/29/compressed-air-energy-storage-becoming-a-leading-energy-storage-technology

You might also like