You are on page 1of 18

Chapter 14

Ocean thermal energy


conversion (OTEC)

14.1 Introduction
The ocean is the world’s largest solar collector. In tropical seas, temperature
differences of about 20−25 C may occur between the warm, solar-absorbing
near-surface water and the cooler 500–1000 m depth ‘deep’ water at and
below the thermocline. Subject to the laws and practicalities of thermody-
namics, heat engines can operate from this temperature difference across
this huge heat store. The term ocean thermal energy conversion (OTEC)
refers to the conversion of some of this thermal energy into useful work
for electricity generation. Given sufficient scale of efficient equipment, elec-
tricity power generation could be sustained day and night at 200 kWe from
2
access to about 1 km of tropical sea, equivalent to 0.07% of the solar input.
Pumping rates are about 6 m3 s−1 of water per MWe electricity production.
The technology for energy extraction is similar to that used for energy effi-
ciency improvement in industry with large flows of heated discharge, but
on a much larger scale.
The attractiveness of OTEC is the seemingly limitless energy of the hotter
surface water in relation to the colder deep water and its potential for
constant, base load, extraction. However, the temperature difference is very
small and so the efficiency of any device for transforming this thermal
energy to mechanical power will also be very small. Even for heating, warm
seawater cannot be spilt on land due to its high salt content. Moreover,
large volumes of seawater need to be pumped, so reducing the net energy
generated and requiring large pipes and heat exchangers.
There have been hundreds of paper studies, and a few experimental
demonstration plants, with the first as far back as 1930. These were mostly
resourced from France (pre-1970s) and then the USA, Japan and Taiwan
in the 1980s, but less activity since then; see Avery and Wu (1994) for a
detailed history. This experience confirmed that the cost per unit of power
output would be large, except perhaps on a very large scale, and led to
other justifications for pumping up the cold, deeper waters, which contain
nutrients and therefore increase surface photosynthesis of phytoplankton
and hence fish population. It now appears that OTEC could be at best a
454 Ocean thermal energy conversion (OTEC)

secondary aspect of systems for deep-water nutrient enrichment for marine


fisheries, for cooling buildings or for desalination (see Section 14.5). Such
integrated technology is called Deep Ocean Water Application (DOWA).

14.2 Principles
Figure 14.1 outlines a system for OTEC. In essence it is a heat engine with
a low boiling point ‘working fluid’, e.g. ammonia, operating between the
‘cold’ temperature Tc of the water pumped up from substantial depth and
the ‘hot’ temperature, Th = Tc + T , of the surface water. The working
fluid circulates in a closed cycle, accepting heat from the warm water and
discharging it to the cold water through heat exchangers. As the fluid
expands, it drives a turbine, which in turn drives an electricity generator.
The working fluid is cooled by the cold water, and the cycle continues.
Alternative ‘open cycle’ systems have seawater as the working fluid, but
this is not recycled but condensed, perhaps for distilled ‘fresh’ water; the
thermodynamic principles of the open cycle are similar to the closed cycle.
In an idealised system with perfect heat exchangers, volume flow Q of
warm water passes into the system at temperature Th and leaves at Tc (the
cold water temperature of lower depths). The power given up from the
warm water in such an ideal system is

P0 = cQT (14.1)

Figure 14.1 Schematic diagram of an OTEC system. A heat engine operates between
the warm water from the ocean surface and the cold water from the
ocean depths.
14.2 Principles 455

where

T = Th − Tc

The second law of thermodynamics dictates that the maximum output of


work energy E1 obtainable from the heat input E0 is

E1 = Carnot E0 (14.2)

Naively forgetting time dependence and the practicalities of heat exchang-


ers, this is usually also given as

P1 = Carnot P0 (14.3)

where

Carnot = T/Th (14.4)

is the efficiency of an ideal Carnot engine operating at an infinitely slow rate


between Th and Tc = Th − T . With T only ∼20 C =20 K, even this ideal
Carnot efficiency is very small, ∼7%. In practice, we cannot wait an infinite
time for ideal thermal processes, so no practical system ever reaches Carnot
efficiencies or a perfect heat exchange. So allowing for temperature drops of
∼5 C across each heat exchanger and for the internal power for pumping, the
efficiency of a real system will be substantially less at about 2–3%. Neverthe-
less these equations suffice to illustrate the promise and limitations of OTEC.
From (14.1)–(14.4) the ideal mechanical output power is

P1 = cQ/Th T2 (14.5)

Example 14.1 Required flow rate


For T = 20C the flow rate required to yield 1.0 MW from an ideal
heat engine is (from (14.5))

106 J s−1 300 K


Q1 = −3 −1
103 kg m 42 × 103 J kg K−1 20 K2
= 018 m3 s−1
−1
= 650 t h

Example 14.1 shows that a substantial flow is required to give a reasonable


output, even at the largest T available in any of the world’s oceans. Such
a system requires large, and therefore expensive, machinery.
456 Ocean thermal energy conversion (OTEC)

Figure 14.2 Seasonal average of temperature difference T between sea surface and a
depth of 1000 m. Zones with T ≥ 20C are most suitable for OTEC. These
zones all lie in the tropics. Source: US Department of Energy.

Since P1 depends quadratically on T , experience shows that only sites


with T ≥ 20C may possibly be economic. Figure 14.2 indicates that such
sites are confined to the tropics, and Figure 14.3 suggests that the cold
water has to come from a depth >∼400 m.
Sites investigated include Hawaii (20N, 160W), Nauru (0S, 166E) and
the Gulf Stream off Florida (30N, 80E). Tropical sites have the added
advantage that both Th and Tc have little seasonal variation, so the potential
output of the system is constant through the year.
14.2 Principles 457

Figure 14.3 Ocean conditions offshore from the island of Nauru, in the Central Pacific
Ocean (0S, 166  E). (a) Water temperature. (b) Cross-section of sea bottom.
The water temperatures are typical of those at good OTEC sites, and the
steeply sloping sea floor allows a land-based system. Data from Tokyo Electric
Power Services Co. Ltd.

Indeed steadiness and independence of the vagaries of weather are major


advantages of OTEC as a renewable source of energy. Its other major
advantages as a possible technology are:

1 At a suitable site, the resource is essentially limited only by the size of


the machinery.
2 The machinery to exploit it economically requires only marginal
improvements in such well-tried engineering devices as heat exchangers
and turbines. No dramatically new or physically impossible devices are
required.

The major disadvantages are cost and scale. Even if the ideal power
P1 of (14.5) was obtainable, the costs per unit output would be large,
but resistances to the flow of heat and to fluid motion reduce the useful
output considerably and therefore increase unit costs. Sections 14.3 and 14.4
estimate the energy losses due to imperfect heat exchangers and pipe friction.
The installed costs of the best experimental OTEC plants (1980s to 1990s)
were as large as $40 000 per kWe of electricity capacity, in comparison with
about $1000 per kWe for conventional generating capacity in remote areas.
However, the theory of Sections 14.2 to 14.4 suggests that even larger
systems would be more economical, which maintains interest in OTEC.
However, a large scale-up in a single step from small demonstration plants
is imprudent engineering and therefore difficult to finance.
Factors increasing the cost of offshore OTEC are maintenance at sea and
submarine cabling, as discussed further in Section 14.5. However, there are
a few especially favourable coastal sites where the sea bed slopes down so
steeply that all the machinery can be placed on dry land. The island of
458 Ocean thermal energy conversion (OTEC)

Figure 14.4 Experimental land–based OTEC plant on Nauru, built by Tokyo Electric
Power Services Company in 1981 for research. It was a ‘closed cycle’ system,
rated at 100 kWe output. On the photograph the vertical framework to
the rear contains the condenser, the nearer large horizontal cylinder is the
evaporator, the turbine house is at the left, the cold water pipe runs out
to sea (in the background), and cylinders in the foreground contain spare
working fluid.

Nauru in the South Pacific has such topography. Figure 14.3 shows a section
of the sea bottom there, and Figure 14.4 is a photograph of an experimen-
tal OTEC installation on the shore. Experience showed (i) the beach and
submarine pipes must be buried or fixed extremely well to survive wave
current forces, (ii) biofouling could be mitigated by 24-hourly pulses of
chlorination, and (iii) in the pipes, both thermal losses and friction decreased
efficiency significantly.

14.3 Heat exchangers


These need to be relatively large to provide sufficient area for heat transfer
at low temperature difference, and are therefore expensive (perhaps 50% of
total costs). In calculating the ideal output power P1 as calculated in (14.5),
we have assumed perfect heat transfer between the ocean waters and the
working fluid. In practice, there is significant thermal resistance, even with
the best available heat exchangers and with chemical ‘cleaning’ to lessen
internal biofouling.
14.3 Heat exchangers 459

14.3.1 General analysis


A heat exchanger transfers heat from one fluid to another, while keeping the
fluids apart. Many different designs are described in engineering handbooks,
but a typical and common type is the shell-and-tube design (Figure 14.5).
Water flows one way through the tubes while the working fluid flows
through the shell around the tubes.
Figure 14.6 shows some of the resistances to heat transfer. The most
fundamental of these arises from the relatively small thermal conductivity
of water. As in Section 3.4, one can think of heat being carried by blobs
of water to within a fraction of a millimetre of the metal surface, but, even
with clean surfaces, the last transfer from liquid to solid has to be by pure
conduction through effectively still water. Similarly the heat flow through
both the metal and the adhering scum and biological growth is by pure
conduction. A temperature difference T is required to drive the heat flow
across these conductive resistances.
Let Pwf be the heat flow from water (w) to working fluid (f). Then
(14.6)
Pwf = T/Rwf
where Rwf is the thermal resistance between water and fluid. If it is assumed
that there will be a similar temperature drop T in the other heat exchanger,

Figure 14.5 Shell-and-tube heat exchanger (cut-away view).

Figure 14.6 Resistances to heat flow across a heat exchanger wall.


460 Ocean thermal energy conversion (OTEC)

the temperature difference actually available to drive the heat engine is not
T but

2 T = T − 2T (14.7)

With an idealised Carnot engine the mechanical power output would be


 
T − 2T T
P2 = (14.8)
Th Rwf
Equation (14.8) implies that T/Rwf should be large to increase output
power. Yet T must be small to obtain maximum engine efficiency, so it is
crucial to minimise the transfer resistance Rwf by making the heat exchanger
as efficient as possible. Therefore the tubes must be made of metal (good
conductor) and there must be many of them, perhaps hundreds, to provide
a large total surface area. Other refinements may include fins or porous
surfaces on the tubes, and baffles within the flow. With such an elaborate
construction, it is not surprising that the heat exchangers constitute one of
the major expenses of an OTEC system. This is the more so since the tube
material has to be resistant to corrosion by seawater and the working fluid,
all joints must be leakproof and all pipes capable of internal cleaning.
The overall thermal resistance can be analysed in terms of the thermal
resistivity of unit area rwf and the total wall area Awf , as in Section 3.6:

Rwf = rwf /Awf (14.9)

Much of the development work in OTEC concerns improvements in the


design of existing heat exchangers. The aim is to decrease rwf , and thereby
decrease the area Awf . Having smaller heat exchangers with less metal can
lead to substantial cost reductions. Values for rwf of 3 × 10−4 m2 KW−1 (i.e.
h = 1/r = 3000 W m−2 K−1 ) can be obtained by the best of existing technology.
The flow rate required through the heat exchanger is determined by
the power Pwf removed from the water, and by the heat transfers and
temperatures involved. These are indicated in Figure 14.7, which shows a
counterflow heat exchanger on each side of the working fluid circuit. At each
point along the heat exchanger, the temperature difference between the work-
ing fluid and the water is T . Thus the hottest point in the working fluid is at
in
Thf = Thw − T

and the coldest is at

Tcf = Tcw
in
+ T

Therefore the power given up by the hot water is


 
in out
Pwf = cQ Thw − Thw (14.10)
14.3 Heat exchangers 461

Figure 14.7 Temperatures and heat flows in the OTEC system of Example 14.2. The
in in
other quantities are calculated from Tcw , Thw , T, P2 .

with the temperature drop


in out
Thw − Thw = T − 2 T (14.11)

14.3.2 Size

Example 14.2 Heat exchanger dimensions


Find a set of working dimensions for a shell-and-tube heat exchanger
suitable for an OTEC system set to produce 1 MW. Assume a Carnot
cycle for the working fluid, but allow for temperature reductions in
non-perfect heat exchangers.
Assume rwf = 3 × 10−4 m2 K W−1 , T = 20 C, T = 4 C, etc. as in
Figure 14.7.
Solution

1 Surface area
From (14.9),

Awf = rwf /Rwf

From (14.8),

P2 Th
1/Rwf =
T − 2 T  T
462 Ocean thermal energy conversion (OTEC)

so
1 × 106 W300 K3 × 10−4 m2 K W−1 
Awf =
20 − 8 K4 K
= 19 × 103 m2

This is a very large area of transfer surface.


2 Flow rate
For the parameters of Figure 14.7,

21 − 9C 12
carnot = =
273 + 21 K 294
Pwf = P2 /carnot = 1 MW294/12 = 25 MW

Therefore, from (14.10), and (14.11), the flow rate is

25 × 106 W
Q= −3 −1
103 kg m 42 × 103 J K−1 kg 12 K
= 050 m3 s−1

3 Thermal resistance of the boundary layers


We suppose that each fluid boundary layer of Figure 14.6 contributes
about half of rwf . In particular, assume that the thermal resistivity
of the boundary layer (of water) on the inside of the pipe is given by

rv = 15 × 10−4 m2 K W−1

Let d be the diameter of each tube in the heat exchanger. The


convective heat transfer to the inside wall of a smooth tube is
given by (C.14), see Section 3.4:

 = 002708  033

By definition of the Nusselt number,  = d/rv k. Thus the


Reynolds number in each tube is

 =
d/0027rv k 033  125
=
002706 W m−1 K−1 70033 −125 d/rv 125
= ad125

where a = 467 × 106 m−125 and the properties of water are from
Appendix B.
14.3 Heat exchangers 463

4 Diameter of tube
As an initial estimate, suppose d = 002 m. Then  = 35 × 104 .
Hence, flow speed in each tube is

u=
d
35 × 104 10 × 10−6 m2 s−1 
= = 17 m s−1
002 m

Since the total flow through n tubes is Q = nu d2 /4 the number


of tubes required is

050 m3 s−1 4


n=
17 m s−1 314001 m2
= 3600

5 Length of tubes
To make up the required transfer area A = n dl, each tube must
have length

19 × 103 m2 
l= = 32 m
3600 002 m

This example makes it clear that large heat exchangers, with substantial
construction costs, are required for OTEC systems. Indeed the example
underestimates the size involved because it does not allow for imperfections
in the heat engines etc., which increase the required Q to achieve the
same power output. Also the example assumes that the pipe is clean and
smooth.
14.3.3 Biofouling
The inside of the pipe is vulnerable to encrustation by marine organisms, which
will increase the resistance to heat flow (Figure 14.6), and thereby reduce the
performance. Such biofouling is one of the major problems in OTEC design,
since increasing the surface area available for heat transfer also increases the
opportunity for organisms to attach themselves. Among the methods tried to
keep this fouling under control are mechanical cleaning by continual circula-
tion of close fitting balls and chemical cleaning by additives to the water.
The effect of all these complications is that the need for cost saving
encourages the use of components working at less than optimal perfor-
mance, e.g. undersized heat exchangers.
464 Ocean thermal energy conversion (OTEC)

14.4 Pumping requirements


Work is required to move large quantities of hot water, cold water and
working fluid around the system against friction. This will have to be sup-
plied from the gross power output of the OTEC system, i.e. it constitutes yet
another loss of energy from the ambient flow P0 . Example 14.3 shows how
the work may be estimated numerically using the methods of Section 2.6,
although analytic calculations are difficult. The effect of cooling the water
in the hydrostatic ‘circuit’ is small, but does encourage circulation.

Example 14.3 Friction in the cold water pipe


The OTEC system of Example 14.2 (Figure 14.7) with P2 =
1 MW T = 20 C has a cold water pipe with L = 1000 m, diameter
D = 1 m. Calculate the power required to pump water up the pipe.
Solution
The mean speed is

u = Q/A
050 m3 s−1 
= = 63 m s−1
05 m 2

Therefore the Reynolds number is

uD
R=
v
063 m s−1 1 m
= = 63 × 105
10 × 10−6 m2 s−1 

In practice, many varieties of marine organisms brought up from the


depths will adhere to the pipe, giving an equivalent roughness height
 ∼ 20 mm, i.e. /D = 002. Thus, from Figure 2.6, the pipe friction
coefficient is f = 0012
From (2.14), the head loss is

Hf = 2fLu2 /Dg = 10 m

To overcome this requires the same power as to lift a mass Q per


second through a height Hf , i.e.

Pf = QgHf = 47 kW
14.5 Other practical considerations 465

From Example 14.3, we see that the cold water pipe can be built large
enough to avoid major friction problems. However, because the head loss
varies as diameter−5 (See problem 2.6), friction loss can become apprecia-
ble in the smaller piping between the cold water pipe and the heat exchanger,
and in the heat exchanger itself. Indeed, because the same turbulence carries
both heat and momentum from the heat exchanger surfaces, all attempts
to increase heat transfer by increasing the surface area necessarily increase
fluid friction in the heat exchangers.
In addition, the flow rate required in practice to yield a given output
power is greater than that calculated in Example 14.2, because a real heat
engine is less efficient than a Carnot engine in converting the input heat
into work. This increases the power lost to fluid friction. Fouling of the
heat exchanger tubes makes the situation worse, both by further raising
the Q required to yield a certain power output, and by decreasing the tube
diameter. As a result, in some systems over 50% of the input power may be
lost to fluid friction. Power used by the pumps themselves is another ‘loss’
from the output power.

14.5 Other practical considerations


The calculations of the previous sections confirm that there are no fun-
damental thermodynamic difficulties that prevent an OTEC system from
working successfully. Although there remain a number of practical, engi-
neering and environmental difficulties, we shall see that none of these
appears insuperable from a technical point of view.

14.5.1 The platform


American designers drew up conceptual plans for large systems, generating
electricity at about 400 MWe , based on a large floating offshore platform,
similar to those used in oil drilling. Since such a platform would be heavy
and unwieldy, there would be a major problem in connecting it to the cold
water pipe (CWP), because of the stresses from surface waves and currents.
One response to this problem is to make the platform neutrally buoyant
and moor it underwater (Figure 14.8) thereby avoiding the major stresses
at the surface.

14.5.2 Construction of the cold water pipe


The pipe is subject to many forces in addition to the stresses at the con-
nection. These include drag by currents, oscillating forces due to vortex
shedding, forces due to harmonic motion of the platform, forces due to
drift of the platform and the dead weight of the pipe itself. It is debat-
able whether a rigid, e.g. steel, or flexible material, e.g. polythene, would
466 Ocean thermal energy conversion (OTEC)

Figure 14.8 Underwater platform for 400 MWe systems; proposed by Lockheed for
the US Department of Energy. The platform can be moored in position
in any depth of water.

withstand these forces better. In addition, there are substantial difficulties


involved in assembling and positioning the pipe. Some engineers favour
bringing out a prefabricated pipe and slowly sinking it into place; however,
transporting an object several meters in diameter and perhaps a kilometre
long is difficult. Premature failure of the CWP, e.g. from storm damage,
caused the failure of several demonstration projects.

14.5.3 Link to the shore


High voltage, large power, submarine cables are standard components of
electrical power transmission systems. They are expensive, as with all marine
engineering, but a cable about 50 km long is quite practicable, with power
loss about 0.05% per km for AC and 0.01% per km for DC. There is now
considerable experience with such cables for offshore wind power and for
underwater connections in power-grid networks.
Alternatively it has been suggested that large OTEC plants, which might
be hundreds of kilometres away from energy demand, could use the electric-
ity on board to produce a chemical store of energy, e.g. H2 , Section 16.3.
Land-based systems, like that of Figure 14.4, are possible at certain
favourable locations, where the sea bed slopes sharply downward. Their
main advantage is reduced cost, since the link to shore, assembly and main-
tenance are much simplified. The CWP is also not so subject to stress, since
it rests on the sea bottom; however, it is still vulnerable to storm damage
from wave motion to a depth of about 20 m.
14.5 Other practical considerations 467

14.5.4 The turbine


Even though the turbine has to be large, standard designs can be used. For
example, engines for working across relatively small temperature differences
have been developed and used in Israel in connection with solar ponds,
Section 6.7. As with all practical heat engines, the efficiency will not be
greater than 50% of an ideal Carnot engine with the same heat input to the
working fluid.

14.5.5 Choice of working fluid


There are many common fluids having an appropriate boiling point, e.g.
ammonia, freon or water, but many of these are environmentally unaccept-
able, since leaks increase greenhouse or ozone-depleting gases. By applying
a partial vacuum, i.e. reducing the pressure, the boiling point of water can
be reduced to the temperature of the warm water intake. This is the basis
of the open cycle system, in which the warm seawater itself is used as the
working fluid. Such a system provides not only power but also substantial
quantities of distilled water.

14.5.6 Related technologies


OTEC is one of several possible deep ocean water applications (DOWA)
associated with pumping seawater from depths of at least 100 m. Others are
listed below. All have dimensional scaling factors encouraging large equip-
ment, unlike the modular operation and smaller scale of most renewable
energy options.

a Marine farming. Seawater from the depths below about 500 m is rich in
nutrients, and these may be pumped to the surface, as from an OTEC
plant. This encourages the growth of algae (phytoplankton), which feed
other marine creatures higher up the food chain and so provides a basis
for commercial fish farming.
b Cooling. Deep, cool water pumped to the surface may be used to cool
buildings, tropical horticultural ‘greenhouses’ or engineering plants as
in chemical refineries.
c Fresh water. Flash evaporation of upper surface sea water onto con-
densers cooled by deep water produces ‘distilled’ ‘fresh’ water for
drinking, horticulture, etc. This process may be integrated with solar
distillation.
d CO2 injection. The aim is to absorb CO2 emitted from large-scale fossil
fuel combustion by absorption into surface sea water and pumping to
depth. This is almost the reverse of the technology for the OTEC CWP,
and would be on a very large scale. Environmental impact on the biota
at depth is an issue, as are cost and sustainability.
468 Ocean thermal energy conversion (OTEC)

e Floating industrial complexes. Concepts exist to match the large scale of


OTEC and DOWA with industry on very large, km scale, floating rafts,
e.g. for ammonia and hydrogen production for shipping to land-based
markets. Talk is cheap!

If OTEC, or similar technologies, are ever to become accepted commercially,


it seems inevitable that an integrated set of operations will be used for a
combination of several benefits.

14.6 Environmental impact


The main environmental impacts of OTEC-like technologies relate to:

• small thermodynamic efficiencies of engineering plant which in turn


relate to the relatively small temperature differences of about 25C
between surface and deep water;
• leakage, and likely pollution, from engineering plant, especially of the
working fluids and antifouling chemicals;
• consequent large volumes of pumped marine water;
• forced mixing of deep nutrient-rich (nitrate, phosphate and silicate)
water with upper, solar irradiated, water;
• location of engineering plant.

Local pollution must always be avoided. Otherwise none of these impacts


appear to be particularly grave on a global scale unless very large numbers of
OTEC systems are deployed. The hypothetical location of very many OTEC
plants, say 1000 stations of 200 MW e each in the Gulf of Mexico, has been
calculated to reduce surface sea temperature by 03C. Such a reduction,
even at such an unlikely scale, is not considered physically significant.
Of more significance locally would be impacts from onshore OTEC or
DOWA engineering plant with local waters, including local circulation and
currents.
The total biological effects of releasing large quantities of cool, nutrient-
rich water into the warmer surface environment are not fully known. The
effects may or may not be desirable, and have to be estimated from small-
scale trials and computer modelling. Large deployment of OTEC plant, say
100 stations at 10 km separation, would cause the upwelling of nitrate to
a concentration found naturally off Peru, where fish populations are much
increased. Consequently immediate impacts need not all be negative to
mankind, and certainly the prospect of enriching fisheries with deep water
nutrients is considered as potentially positive. As cold, deep water reaches
the ocean surface, a proportion of dissolved CO2 passes into the atmosphere.
−1
If 50% of the excess CO2 is emitted, the rate would be about 01 kg kWe ,
Bibliography 469

−1
as compared with about 08 kg kWe from electricity generation by fossil
fuel. Only if the OTEC energy produced is used to abate the use of fossil
fuels are global emissions of CO2 reduced.
The social impacts of OTEC would be similar to those of running an
offshore oil rig or an onshore power station, i.e. minimal.

Problems
14.1 Calculate the dimensions of a shell-and-tube heat exchanger to pro-
duce an output power P2 = 10 MW. Assume rv = 3 × 10−4 m2 K W−1 ,
T = 4 C. and tube diameter D = 5 cm.
Hint: Follow Example 14.2.
14.2 Calculate the power lost to fluid friction in the heat exchanger of
Example 14.2.
14.3 Heat engine for maximum power. As shown in textbooks of thermo-
dynamics, no heat engine could be more efficient than the ideal con-
cept of the Carnot engine. Working between temperatures Th and Tc ,
its efficiency is

Carnot T = Th − Tc /Th

However, the power output from a Carnot engine is zero. Why?


Use (14.8) to show that the engine which produces the greatest
power, for constant thermal resistance of pipe, has T = 1/4T , i.e.
it ‘throws away’ half the input temperature difference. What is the
efficiency of this engine as an energy converter compared with an
ideal Carnot engine?
14.4 If P ∝ T 2 /Th (14.4), calculate the rate of change of efficiency with
respect to temperature difference T . What is the percentage improve-
ment in power production if T increases from 20 to 21C?

Bibliography

Monographs
Avery, W.H. and Wu, C. (1994) Renewable Energy from the Ocean – A Guide to
OTEC, Oxford University Press (John Hopkins University series). A substantial
and authoritative study of the science, engineering and history of OTEC.
Ramesh, R., Udayakumar, K. and Anandakrishnan, M. (1997) Renewable Energy
Technologies: Ocean Thermal Energy and Other Sustainable Energy Options,
Narosa Publishing, London and Delhi. Collection of optimistic papers from a
conference on OTEC in Tamil Naidu, India.
470 Ocean thermal energy conversion (OTEC)

Articles
d’Arsonval, Jacques (1881) Revue Scientifique, 17, pp. 370–372. Perhaps the ear-
liest published reference to the potential of OTEC.
Gauthier, M., Golman, L. and Lennard, D. (2000) Ocean Thermal Energy Con-
version (OTEC) and Deep Water Applications (DOWA) – market opportunities
for European Industry, in Proc. Euro. Conf. New and Renewable Technologies
for Sustainable Development, Madeira, June 2000. Excellent review of working
plant since the 1930’s to 2000, with future industrial market potential.
Johnson, F.A. (1992) Closed cycle thermal energy conversion, in Seymour, R.J.
(ed.), Ocean Energy Recovery: The State of the Art, American Society of
Civil Engineers (1992). Useful summary of thermodynamics, economics and
history.
Masutani, S.M. and Takahashi, P.K. (1999) Ocean Thermal Energy Conversion, in
J.G. Webster (ed.) Encyclopaedia of Electrical and Electronics Engineering, 18,
pp. 93–103, Wiley. Authoritative summary.
McGowan, J.G. (1976) Ocean thermal energy conversion – a significant solar
resources, Solar Energy, 18, pp. 81–92. Reviewed US design philosophy at a
historically important time.
Ravidran, M. (1999) Indian 1 MW Floating Plant: An overview, in Proc. IOA ‘99
Conf., IMARI, Japan.
UN (1984) A guide to Ocean Thermal Energy Conversion for Developing Countries,
United Nations Publications, New York.
Wick, G.I. and Schmidt, W.R. (1981) (eds) Harvesting Ocean Energy, United
Nations, Paris.
Zener, C. (1974) Solar sea power, in Physics and the Energy Problem – 1974,
American Institute of Physics Conference Proceedings no. 19, pp. 412–419.
Useful for heat exchanger thermodynamics. Whole volume makes interesting
reading.

Thermodynamics of real engines


Curzon, F.L. and Ahlborn, B. (1975) Efficiency of a Carnot engine at maximum
power output, Amer. J. Phys., 43, pp. 22–24.

You might also like