You are on page 1of 32

652 Int. J. Exergy, Vol. 5, Nos.

5/6, 2008

Energy analysis of Brayton combined cycles

A.L. Polyzakis
Prefecture of Western Macedonia,
Department of Public Works Administration,
Epivaton 39, Ptolemaida, 50200, Greece
E-mail: apolyzakis@yahoo.gr

C. Koroneos*, G. Xydis and A. Malkogianni


Laboratory of Heat Transfer and Environmental Engineering,
Aristotle University of Thessaloniki,
P.O. Box 483, GR. 54124, Thessaloniki, Greece
E-mail: koroneos@aix.meng.auth.gr
E-mail: xydis@aix.meng.auth.gr
E-mail: arita_m20@yahoo.gr
*Corresponding author

Abstract: Economic and environmental considerations dictate the need for


improved fuel utilisation in thermal power stations. To meet this goal, a design
project was undertaken to assess the suitability for Combined Cycle application
of different Gas Turbine (GT) cycles: Simple, Intercooled, Reheated,
Intercooled/ Reheated, Intercooled/Reheated/Regenerative cycle. Comparative
figures for efficiency and performance were obtained. Exergy analysis will lead
to an optimum plant. The proposed Combined Cycle plant would produce
300MW of power (200MW GT cycle and 100MW steam cycle). The reheated
GT is the most desirable exergetically, mainly due to its high Exhaust Gas
Temperature which results in high steam cycle thermal efficiency.

Keywords: exergy; GT; gas turbine; steam; simple; IC; intercooled; RH;
reheated; regenerative cycle.

Reference to this paper should be made as follows: Polyzakisz, A.L.,


Koroneos, C., Xydis, G. and Malkogianni, A. (2008) ‘Energy analysis of
Brayton combined cycles’, Int. J. Exergy, Vol. 5, Nos. 5/6, pp.652–683.

Biographical notes: Apostolis L. Polyzakis works at the Prefecture of Western


Macedonia and is also a Lecturer at the Technical University of Western
Macedonia, Greece. His PhD awarded by Cranfield University deals with the
development of tri-generation power units.

Christopher Koroneos is a Chemical Engineer. He received his PhD, MSc and


BSc, all from Columbia University in the City of New York in USA, where he
also taught for eight years. Presently, he is teaching at the Aristotle University
of Thessaloniki and the University of Western Macedonia. He is also a Visiting
Professor at the National Technical University of Athens in the graduate
programme ‘Environment and Development’.

Copyright © 2008 Inderscience Enterprises Ltd.


Energy analysis of Brayton combined cycles 653

George Xydis is working as a Scientific Personnel in the Unit of Environmental


Science and Technology (UEST) at the National University of Athens (NTUA),
Greece. He is a PhD candidate at the School of Chemical Engineers of NTUA.
He received his Mechanical Engineering Degree from the Aristotle University
of Thessaloniki.

Areti Malkogianni is a PhD student at the University of Western Macedonia.


Her PhD concerns the development of cogeneration power units using biofuels.

1 Introduction

Power plants where Gas Turbines (GTs), operating in open cycles, are combined with
steam cycles are particularly interesting for their high conversion efficiency.
The combination of the two kinds of cycles is possible due to the high temperature
difference between the gas working cycle and the low temperature heat dissipated in the
steam turbine. A great proportion of the energy remaining in the GT exhaust gases can be
readily converted into useful electricity through the use of a Heat Recovery Steam
Generator (HRSG) integrating both, the Brayton cycle and the Rankine or steam cycle.
Gas and steam turbine Combined Cycles Power Plants (CCPPs) have become the major
new force for power generation throughout the world (Matta, 2000; Heppenstall, 1998).
Operating on natural gas CC power plants provide the most efficient method of
power generation with thermal efficiencies reaching 60%. The state-of-the-art GT
for combined cycle is the General Electric Frame 9 (H technology) machine, which is
designed for use in combined cycle only. This machine has a pressure ratio of 23 which is
above the range just mentioned. This is due to the high firing temperature of 1,703 K.
It is claimed that this CC arrangement gives a thermal efficiency of 60% (Asea Brown
Boveri, 2004; GE Energy, 2004; GE Power Systems, 2004).
CCGTs generally pose much fewer problems in terms of site availability since they
require less space; lack the need for transport of fuel by road and rail; require much
smaller fuel handling/waste disposal facilities; are flexible in both construction and
operation, while it can also easily be employed on a modular basis to build power plants
of any size but the most important argument in favour of CCPPs is their level of
emissions, an issue gaining importance every year due to acclaimed contributions to
environmental effects such as global warming and depletion of the ozone layer. There is
another serious contender for power production, namely the conventional coal
combustion plant. Although not as efficient (currently around 45%), in terms of
electricity cost per unit capital cost it can still beat CCPPs, especially in areas where
coal is abundant and cheap. However, to get the emissions of coal-fired plants down to
the levels required by legislation nowadays, very sophisticated gas cleaning processes
must be installed, with severe penalties in efficiency and cost. These measures are
unnecessary when burning natural gas, because it is a very clean fuel. For the same
reason, it allows waste heat recovery to a lower temperature because its combustion
products do not contain any sulphurous components which might form sulphuric acid and
other corrosive compounds. In recent years, therefore, the gas-fired combined cycle plant
has become the most environmentally acceptable way to produce electric power, and will
be the power generation option of choice in the future (Poulikas, 2004; Pilavachi, 2002;
Mahi et al., 1994).
654 A.L. Polyzakisz et al.

To obtain the maximum benefits from a CC power plant it must be designed with
consideration for site specific technical, environmental and economic considerations.
Of paramount importance is the selection of GT, with decisions that affect the rest of the
plant. Having in mind the above outlined importance of finding the way to increase the
efficiency of thermal power generation, the main purpose of this work is to investigate
how this can be achieved in gas and steam combined cycle plants through improvements
in the GT cycle configuration and through the optimisation of the steam bottoming cycle.
The target of the analysis is to determine the optimum technical layout of an advanced
combined gas and steam power plant in which a compromise between fuel consumption,
installation investment and reliability can be guaranteed.
Numerous investigations, as well as experience with actual plants, have showed that
the split of power output between the GT and the steam turbine of a Combined Cycle
plant should be approximately 2 : 1, resulting in a GT of 200 MW shaft power output and
a steam turbine of 100 MW, giving a total power output of 300 MW (Polyzakis, 2007).

2 The Gas Turbine (Brayton) cycle

GTs have been used for industrial applications since not too long ago. It was just
around the 1940s, after World War II, when the industrial open cycle became a reality.
GT knowledge gained in the development of aircraft jets was applied to expand the
already well established basic knowledge in axial flow machinery. At that time fuel was
very expensive and wages were relatively low, thus the main target of engineering efforts
was to raise the efficiency of thermal power generation. GTs were introduced in the hope
that they would achieve better figures than steam turbines. In fact, GTs offered a way of
increasing the upper cycle temperature above the steam level because the cycle operated
at much lower pressures (Alabdoadaim, 2006).
Complex cycles were designed and a number of GTs of these kinds were built and
installed. Nevertheless, despite all these efforts, the GTs working under simple cycle
could not yet reach the efficiencies of the steam turbines, especially the ones with
supercritical characteristics (Hans Dieter Baehr, 1999). The temperature that was possible
to achieve was still too close to the figure used for steam turbines, the result being that
the steam turbine’s far lower sensitivity to thermodynamic irreversibility prevented
the GT cycle from clearly gaining the market. As oil prices dropped and wages
and the cost of materials rose, the GT, in its simplest cycle configuration, became quite
well suited for peak load and emergency power generation, since fuel consumption was
therefore more or less irrelevant and the first priority of utility companies was to keep
low investment costs.
The search for increased efficiencies, together with the drastic reduction of pollutant’s
emissions, has never stopped. Achieving better thermal efficiencies while keeping
competitive costs has proved to be a formidable task and has involved, and/or been fed
from, the parallel development of other sciences such as aerodynamics and metallurgy,
which have made possible the use of continuously increased levels of pressure and
temperature, this in turn leading to a substantially improved fuel utilisation and hence to a
lower level of contaminants emitted from thermal power plants. The traditional route of
GT development for gaining higher output and efficiency has been through raising firing
temperatures, accompanied by an increase of mass flow in a larger frame size. However,
in the environmentally conscious market of today there is a clear conflict of interest with
Energy analysis of Brayton combined cycles 655

this approach, since thermal NOx increases exponentially. This is, therefore, the reason
for manufacturers to have chosen the option of increasing power density and mass flow
by significantly augmenting the pressure ratio, combining this with sequential
combustion at relatively lower firing temperatures (Saravanamutoo et al., 2001).
Reheating, or sequential combustion, is not new indeed; it was applied to GTs more
than 40 years ago. In its simplest way, for example, the fired heat recovery boiler of a
combined steam and gas power plant can be depicted as a sequential combustion. Fuel is
added to the hot exhaust gases, in this case to generate more steam, and normally steam
output is modulated by varying the fuel input to the burner. As applied to the GT,
reheating allows the hot gases to pass to a second combustion chamber so that more work
can be extracted as shaft horsepower. In this case increase in pollutant emissions is
marginal, because although there is enough oxygen in the exhaust gases of the first stage
to support combustion, it is not sufficient to add significantly to thermal NOx production.
Manufacturers have called into question the future direction of GT development. They
have always argued that making a more efficient GT was the easier route to a more
efficient combined cycle, instead of going to even higher steam conditions. This is clearly
the aim with the new machines (Najjar, 2001).
For years, the GT has mainly been operated in a simple cycle. Figure 1 shows the
temperature-entropy diagram of the basic isentropic Brayton cycle. It consists of one
stage of compression, a single heat input stage, and simple expansion through a turbine.
When considering a combined cycle including a GT, the Rankine steam cycle is simply
added underneath the Brayton cycle, as shown in Figure 1. The Brayton topping cycle
rejects its heat to the Rankine bottoming cycle to raise steam, instead of directing it to the
atmosphere as in the Simple Cycle (SC) configuration. For obvious reasons, this principle
is called ‘waste heat recovery’ and the steam-raising device substituting the boiler of a
conventional steam plant is called the HRSG. The single pressure combined cycle plant
thermal efficiency is defined by (Kolanowski, 2003):
(work output)CC netGT work + net STwork
ηCC = = . (1)
(heat input)GT fuel energy input into GT

Clearly, the efficiency of the combined gas/steam cycle is higher than that of the GT by
itself. However, where the optimisation of one cycle was relatively simple – i.e., a matter
of raising the mean temperature of heat reception and lowering that of heat rejection-the
optimisation of the CC is a more complicated matter. Το investigate this issue further, the
first four potential (SC, IC, RH, IC/RH) GT topping cycles were analysed for their
suitability for combined cycle application, Figure 1.
Α GT simulation software program called TurboMatch has been used to simulate the
performance of the different cycles. This program has been developed and refined over
many years in the Department of Propulsion, Power, and Automotive Engineering
of Cranfield University (Bolland, 1995). It facilitates Design-Point (DP) and Off Design
(OD) performance calculations of aero and industrial engines of any layout. Βy means
of ‘codewords’ various pre-programmed routines known as ‘bricks’ can be combined to
build a model of the desired GT. Basically, the characteristics of each brick, representing
an actual component of the engine, are determined by several parameters that are relevant
to the function of the component.
TurboMatch uses these data to simulate the (inter-) action of the different components
of the modelled engine. It generates output in terms of delivered engine thrust
656 A.L. Polyzakisz et al.

(when modelling aero engines) or shaft power, Specific Fuel Consumption (SFC), etc.,
with details of individual component performance and of the gas path parameters at
various stations within the machine.

Figure 1 Combined cycle: combination of simple Brayton cycle and Rankine cycle

3 Gas Turbine Design Point and Off Design point calculations procedure

Design Point (DP) calculation procedure -referring to simple cycle 1-shaft engine
(Figure 2) – is shown in Table 1. Every component of the engine is characterised by two
numbers (station vectors), one for the inlet and one for the outlet of the component.
The conditions at the outlet of the first component should be the same as the conditions at
the inlet of the second component etc. (Kim and Hwang, 2006, 2007).

Figure 2 Simple cycle 1-shaft GT layout. ‘Brick’ and ‘station vector’ model

For the DP calculation we consider the ISO Atmosphere conditions: (Ambient pressure:
Pamb = 101.3 kPa, Ambient temperature: Tamb = 288 K).
OD calculation procedure -referring again to simple cycle 1-shaft engine (Figure 2),
is shown also on Table 1. There must be a selection of an engine from the DP
performance. This selection should take into account several constraints and restrictions,
depending on the specific application. Consequently, the calculation procedure for
the selected DP should be carried out and then we proceed with the OD calculation
varying the ambient conditions (Arrieta and Silva Lora, 2005).
The simulation program includes two subroutines. Each subroutine simulates
the OD performance of the engine when one of the following parameters varies:
Energy analysis of Brayton combined cycles 657

• Ambient temperature. The operating temperatures of the engine are the average
temperatures of each month of the region where the engine will operate.
• Altitude. The term altitude is used in order to indicate the simultaneous variation of
both ambient temperature and pressure. The operating temperatures and pressures of
the engine are the average values of each month of the region where the engine will
operate.

Table 1 DP and OD calculation procedure

Station Station
vector DP calculation procedure vector OD calculation procedure
γ c /(γ c −1) γ c /(γ c −1)
 γ −1   γ c −1 
Po1 = Pamb ⋅  1 + c ⋅ M m2  Pood1 = Pamb
od
⋅ 1 + ⋅ M in2 
 2   2 
 γ −1  od  γ c −1 
1 To1 = Tamb ⋅ 1 + c ⋅ M in2  1 Tood
1 = Tamb ⋅  1 + ⋅ M in2 
 2   2 
m 1 = m m2 ⋅ Pood2
m od =
To 2

 DPin loss   DPin loss 


Po 2 = Po1 ⋅ 1 −  Pood2 = Pood
1 ⋅ 1 − 
 100   100 
2 2
To2 = To1 Tood2 = Tood
1

m 2 = m m 1od = m 2od = m od
Po 3 = Po 2 ⋅ Rc Pood8 = Pood
1 ⋅ 1.003 (where 1.003 is

assumed the pressure just before the


8 exit of the exhaust nozzle, so the gasses
can exit from the back of the engine to
3 the atmosphere)
Rcγc −1 γc − 1 Pood8 ⋅ 100
To 3 = To 2 ⋅ + To 2 7 Pood7 =
ηisc 100 − DPexh loss
m 3 = m 5 Tood5 = TET od
Pο4 = Pο3 Tood6 = Tood5 − DTcooling
od

To4 = To3
m od ⋅ Po 6  To 6 
od

Pood6 = ⋅ 
6 m  To 6 
 
Dmcooling = 0 (If ΤΕΤ = Tο5 < 1,300K) Pood6
Rtod =
Pood7
4
DTcooling = 0 (If ΤΕΤ = Tο5 < 1,300K) Pood5 = Pood6

 Dm cooling  Pood5
m 4 = m ⋅  1 −  Pood4 =
 100  5 1 − ( DPccloss /100)
FAR 45 = m f / m 4  DPccloss 
ηccod = ηcc ⋅ 1 − 
 100 
658 A.L. Polyzakisz et al.

Table 1 DP and OD calculation procedure (continued)

Station Station
vector DP calculation procedure vector OD calculation procedure
 DPccloss 
Po 5 = Po 4 ⋅  1 −  4 Pood3 = Pood4
 100 
Tο5 = TET Pood3
Rcod =
Pood2
5
m 4 ⋅ (Cph ⋅ To 5 − Cpc ⋅ To 4 )
m f =  ( R od )(γ c −1) / γ c − 1 
ηcc ⋅ FCV ⋅ 1,000,000) 3
Tood3 = Tood2 ⋅  c + 1
(1,000,000 for the units conversion  ηisodc 
 
to MW)
m 5 = m 4 + m f m 3od = m od
Po6 = Po5  Dm od 
cooling
m 4od = m od ⋅  1 − 
 100 
4 
Tο6 = Tο5 – DTcooling Tood4 = Tood3
6
m 6 = m + m f m 4od ⋅ (C ph ⋅ Tood5 − C p c ⋅ Tood4 )
Qccod =
1,000,000
(1,000,000 for the units conversion
to MW)
Po8 = Pamb ⋅ 1.003 (where 1.003 is
assumed the pressure just before the Qccod
exit of the exhaust nozzle, so the 5 m od
f =
8 FCV od ⋅ ηccod
gasses can exit from the back of the
engine to the atmosphere)
m 8 = m 6 m 5od = m 4od + m od
f

Po 8 ⋅ 100 m od
Po 7 = FAR od
45 =
f
100 − DPexh loss m 4od

  ηh −1
 
   Po 7  ηh   m 6od = m od + m od
To 7 = To 6 ⋅ 1 − 1 −   ⋅ ηist  6
7  Po 6  
f

    

m 7 = m 6 ηisod can be determinate from turbine


t

map
(Palrner, 1995)
 DTexh loss 
8 To8 = To 7 ⋅  1 −  7 m 7od = m 6od
 100 
Performance    P od γ h −1/ γ h  
Tood7 = Tood6 ⋅ 1 − 1 −  ood7   ⋅ ηisodt 
   Po 6  
   
Energy analysis of Brayton combined cycles 659

Table 1 DP and OD calculation procedure (continued)

Station Station
vector DP calculation procedure vector OD calculation procedure
m 2 ⋅ Cpc ⋅ (To 3 − To 2 )
CW =
1,000,000 Tood8 = Tood7
(1,000,000 for the units conversion to
MW)
8
m ⋅ Cph ⋅ (To 6 − To 7 )
TW = 6
1,000,000 m 8od = m 6od
(1,000,000 for the units conversion
to MW)
HI = m f ⋅ FCV Performance
UW = TW – CW m od ⋅ C pc ⋅ (Tood3 − Tood2 )
CW od =
1,000,000
(1,000,000 for the units conversion
to MW)
SW = UW / m m od ⋅ C ph ⋅ (Tood6 − Tood7 )
TWod =
1,000,000
(1,000,000 for the units conversion
to MW)
ηth = UW/HI HIod = Qccod

1 UWod = TWod – CWod


sfc =
ηth ⋅ FCV UW od
SW od =
m od
UW od
ηthod =
HIod
1
sfc od =
ηthod ⋅ FCV od

od
m 8od ⋅ C ph ⋅ (Tood8 − Tamb
od
)
Qout =
1,000,000
(1,000,000 for the units conversion
to MW)

4 The Steam (Rankine) cycle

The steam cycle makes use of the exhaust heat from the GT. This would otherwise be
wasted energy. To optimise the steam cycle as much of the heat should be recovered as
possible. The output performance parameters of the GT were input in a steam cycle
calculation process. GT efficiency, exhaust gas mass flow and temperature, and GT
power output were combined with the appropriate steam data.
There are several steam cycle options which can satisfy different economic and
operational requirements (Kolanowski, 2003; Palrner, 1995; Bolland, 1995).
660 A.L. Polyzakisz et al.

• Single-Pressure, Non-Reheat Heat Recovery Cycle (SP). This is the simplest steam
cycle that can be applied in a CC and it has been used extensively. It is shown in
Figure 1. This cycle has an unfired HRSG with finned tube superheater, evaporator
and economiser sections. Energy is recovered from the exhaust gas by convective
heat transfer. Its efficiency is not the maximum but it is the lower installed cost CC
cycle and gives a high stack temperature, so it is appropriate when fuel is
inexpensive, when applied in peak load service or when burning fuels with high
sulphur content. It is therefore preferred when the exhaust GT temperature is
moderate.
• Multiple-Pressure, Non-Reheat Heat Recovery (MP) cycle. In this case the steam
turbine has two or three steam admissions, one for high pressure, and the others for
intermediate and low pressure steam. These cycles achieve better efficiency than the
single-pressure systems, but their installed cost is higher. Their “high investment-low
operational cost” characteristic makes them preferable for base load or when the fuel
is expensive. This cycle is similar to the single pressure cycle with the addition of the
low pressure section. The efficiency is improved because additional heat transfer can
be achieved in the HRSG.
• Multiple-Pressure, Reheat Recovery (MP/RH) cycle. In this case all the steam from
the high pressure cycle is extracted after a first expansion in the turbine and brought
back to the HRSG where it is RH. A second expansion in the turbine follows to this
reheating process. In multiple pressure systems the low pressure stages take care of
the energy remaining from the high pressure stage. This cycle is suited for high
EGTs and provides the highest thermal efficiency.
Note that supplementary firing is not used in power generation but only in the process
plant. It can be shown that the pressure ratio for the GT which will give it the optimum
thermal efficiency is far higher than the pressure ratio which will give optimum thermal
efficiency in the combined cycle.
The single pressure combined cycle is limited in its performance by its heat recovery
capability. The best possible single pressure cycle will have a stack temperature (Tstack)
between 423 K and 453 K. The stack temperature should be as low as possible to
maximise the combined cycle efficiency (Shin and Jeon, 2002). According to Pasha and
Jolly (1995) a heat recovery system with multiple pressures is used, so that the high
pressure system recovers heat at high temperatures and the low pressure system recovers
heat at low temperatures. This has the effect of reducing the stack temperature to as low
as 343 K or 383 K and this give considerable gains in efficiency. The single pressure
level combined cycles are the simplest form of combined cycle. Even with improvements
made arriving at more complex systems they still remain in the market place.
They provide essential understanding to the concepts involved within HRSG which can
then be applied to multi-pressure levels. The single pressure combined cycle is
intrinsically limited in its performance by the heat recovery capability of the single
pressure level system. It can be seen that the stack temperature is limited by: GT
temperature exhaust, Superheat steam temperature, Position and temperature difference
of the pinch point, Temperature of the feedwater (Pasha and Jolly, 1995; Rufli, 1987).
The HRSG is the link between the two cycles. This is where heat exchange between
the hot gases of the GT exhaust and the water/steam of the steam cycle takes place.
A single pressure HRSG separated into components which perform successive tasks to
Energy analysis of Brayton combined cycles 661

produce superheated steam: the heating of liquid water is done in a heat exchanger called
the economiser; the vaporisation is done in a heat exchanger called the evaporator; the
superheated steam is produced in the superheater; the steam drum separates water from
steam. Steam cycles with multiple pressures have more of these basic components in the
HRSG.
The steam turbine produces power by expanding the superheated steam down to
the condenser pressure. The pressure difference across the steam turbine is great with a
condenser pressure around 3 kPa and boiler pressure of up to 17,000 kPa. Increasing this
pressure always increases cycle efficiency but it is difficult to control steam cycles above
17,000 kPa since it is difficult to separate the liquid and vapour phases near the critical
point. The higher the pressure, the more complex and expensive is the HRSG.
For subcritical boilers maximum pressures are 16,000 kPa and maximum steam
temperature 833 K (Bolland, 1995). Using reheat it is possible to increase the steam
pressure to above 22,000 kPa with out any problems in steam quality. The efficiency is
increased in this way by up to 3%.
The condenser is a heat exchanger which condenses the steam/water mixture from the
turbine exhaust into liquid water so that the water can be pumped back to the HRSG.
The condenser pressure varies according to the type of cooling system from 2.5 kPa up to
12 kPa. In general, the condenser pressure should be as low as possible. The quality and
nature of the cooling medium available imposes a minimum limit for the condenser
pressure. There are various methods of extracting the heat of condensation (cooling
system) from the condenser: direct river or sea water cooling – this is the simplest method
but can give fouling and corrosion problems; cooling tower – requires high and costly
tower; district heating. Increasing the superheat temperature increases the average
temperature at which heat is input and therefore increases cycle efficiency.
This temperature should be as high as possible but it has to be limited to 813 K which is a
common creep limit for the ferritic steels which the turbine blades are made from (steam
turbines have uncooled blades). If the steam at the outlet from the turbine contains too
many water droplets then there will be erosion of the turbine blades. The dryness fraction
of the steam should, therefore, be at least 0.88 to avoid this. This must be high enough to
prevent condensation on the HRSG tubes which would lead to corrosion problems. Pinch
Point is the temperature difference between the hot exhaust gas and the superheated
steam (293–303 K). Approach Point is the difference between the saturation temperature
in the evaporator and the water temperature at the exit from the economiser. This
temperature difference prevents steam appearing in the economiser and can be up to
283 K (Khaliq and Kaushik, 2004; Tinsley, 1995; Leenhouts, 1995; Velez, 1995).
It is clear that to achieve the highest Combined Cycle efficiency we need to maximise
the GT thermal efficiency and the steam turbine specific work output from the fixed GT
exhaust energy. The only way to improve the Combined Cycle for a given GT exhaust
flow is to minimise the exergy losses in the boiler and extract the maximum amount of
energy from the gas stream. In real-life applications, the steam cycle is always matched to
the GT as well, not the other way around. GTs are ‘off-the-shelf’ equipment, around
which the steam plant is designed and optimised as site conditions allow.
Α single pressure non-reheat, unfired cycle steam cycle was simulated, the main
outcomes of it being the steam power output, steam cycle efficiency and overall
combined cycle efficiency. The single pressure level combined cycles are the simplest
form of combined cycle. Even with improvements made arriving at more complex
662 A.L. Polyzakisz et al.

systems they still remain in the market place. The following values were adopted for the
steam calculations.
Evaporator pinch point 30 Κ; Minimum stack gas temperature 160 K (Stack
temperature typical value: Single pressure Rankine cycle: 150–180 K, Dual pressure
Rankine cycle: 100–130 K) (Kolanowski, 2003); Steam turbine constant isentropic
efficiency 0.85; Maximum superheated steam temperature 923 K and minimum EGT
823 K; Maximum boiler pressure 16,000 kPa; Minimum condenser pressure 3 kPa;
Approach point 277 K. The dryness fraction of the steam at exit of the steam turbine is an
important parameter from a maintenance point of view. When the dryness fraction is too
low, generally defined as below 0.87 (87% dryness or 13% wetness), the last stage
turbine blades suffer excessive erosion which may lead to reduced efficiency and maybe
even mechanical failure of the turbine.
An average constant value of specific heat (CP) of the air was adopted, in accordance
with the following formula:

CP = 1020 + 420 ⋅ e −800/Tm (2)

in which Tm is the mean gas temperature through the boiler in K, given by:
EGT + Tstack
Tm = . (3)
2
It was assumed that fresh and good quality river water was available for cooling
purposes. Hence, an open condensing cycle was supposed and a very low condenser
pressure was used.
For performance of the steam and combined cycle optimisation a computer model is
used, so that changing parameters could be investigated. Steam performance calculations
can be calculated from the input of data taken from existing engines or via the GT engine
module of the TurboMatch. The steam plant model is based οn the conservation of
energy through the boiler. Steam mass flow can be calculated from an energy balance
across the evaporator and superheater section. With steam mass flow calculated and
steam turbine efficiencies known, the power output can be calculated (Rice, 1986).

5 Simple Cycle (SC)

The configuration of the engine is shown in Figure 2. The DP performance specifications


of the engine follow those shown on Table 2, and they are similar to those given by
the most important manufacturers.
The results from DP calculations are shown in Figure 3. If the engine is operating in
open cycle from this plot of SFC vs. Specific Power (SP) it would be possible to select a
suitable engine for a particular application.
OD performance is carried out to assess the engines’ performance over the complete
range of operating conditions which the engine could experience. These include varying
the ambient temperature conditions (253, 273, 288, 313 K), altitude (500, 1,500, 2,500 m)
and reducing ΤΕΤ to assess the engines performance at part load.
Energy analysis of Brayton combined cycles 663

Table 2 GT’s specifications (Design Point, DP)

Intercooled Reheated Interc./ Interc./Reh./R


Simple Cycle Cycle Cycle Reh. Cycle egen Cycle
Ambient conditions 288 K 288 K 288 K 288 K 288 K
101 kPa 101 kPa 101 kPa 101 kPa 101 kPa
Intake efficiency 100% 100% 100% 100% 100%
Compres. Isent. effic. 86% 86% 86% 86% 86%
Fuel (NG) 43 MJ/kg 43 MJ/kg 43 MJ/kg 43 MJ/kg 43 MJ/kg
Combustion effic. 99% 99% 99% 99% 99%
Burner Pres. loss 5% 5% 5% 5% 5%
Turbine Isent. effic. 87% 87% 87% 87% 87%
TET (HPT, LPT &PT) 1,561 K 1,561 K 1,561 K 1,508 K* 1,561 K
Turbine cooling air 25% 25% 15%, 10% 15%, 10% 15%, 10%
Exhaust pressure 102 kPa 102 kPa 102 kPa 102 kPa 102 kPa
Intercooler effic. – 99% – 99% 99%
Interc. pres. loss – 0% – 0% 0%
Heat exch. effic. – – – – 85%
*Best performance temperature.

Figure 3 Simple cycle Design Point performance

Figures 4 and 5 show the variation of power with ΤΕΤ for varying ambient temperature
and altitude respectively. It can be seen that power output decreases as ΤΕΤ is reduced as
would be expected. Varying the ambient temperature affects the power output. As the
ambient temperature increases the air density falls, hence for a given ΤΕΤ the mass flow
through the engine is reduced. As a consequence, the engine output power is lower.
Furthermore, when the ambient temperature increases, the cycle temperature ratio
(TET/Tamb) decreases and as a result the power output decreases too. Αn increase in
altitude gives rise to the same result as can be seen in Figure 7. It is worth noting that
these figures indicate a continued increase in power with reduced ambient temperatures
and increased ΤΕΤ. In the real situation this would not be permitted due to stress
limitation. Rotors and disks will be mechanically designed to withstand momentary
overspeed of up to 110% (Fielden, 1995).
664 A.L. Polyzakisz et al.

Figure 4 Power vs. TET varying Ta (SC)

Figure 5 Power vs. TET varying altitude (SC)

The variation of ηGT vs. ΤΕΤ for varying ambient temperature and altitude ηGT is shown
in Figure 6 and 7. It can be seen that efficiency reduces rapidly with a reduction in ΤΕΤ.
Thermal efficiency equates to useful work output divided by heat input. Work output is
found to decrease more rapidly than the reduction in heat input, therefore thermal
efficiency decreases.
However, when steam and gas cycles are combined to ascertain the best combination,
a suitable steam cycle must also be considered, (Table 3), it can be seen that the optimum
pressure ratio (RC) for the CC occurs in the region 16 (Polyzakis et al., 2007).

Figure 6 Efficiency vs. TET varying Tamb (SC)


Energy analysis of Brayton combined cycles 665

Figure 7 Efficiency vs. TET varying altitude (SC)

Table 3 Selection of optimum simple cycle GT (TET = 1,561 K)

SPGT Exh.flow St.flow XACT ηST PSTEAM CCPP ηCCPP


RC,TOTAL ηGT (%) (J/kg) (kg/s) EGT (K) (kg/s) (%) (%) (MW) (MW) (%)
12 31.67 284,900 716.6 833.4 98.97 12.8 36.0 117.6 317.6 50.30
13 32.24 285,714 714.3 822.0 96.98 12.9 35.5 112.1 312.1 50.31
14 32.69 285,714 714.2 811.8 94.49 12.9 35.2 108.3 308.3 50.39
15 33.09 285,307 715.0 802.5 92.24 12.8 35.0 104.9 304.9 50.45
16 33.46 284,697 716.4 793.7 91.16 12.8 34.5 101.1 301.1 50.47
17 33.76 283,688 718.7 785.7 89.12 12.8 34.4 98.8 298.8 50.44
18 34.04 282,486 721.6 778.2 88.21 12.9 34.1 96.2 296.2 50.41
19 34.25 280,899 725.5 771.4 87.69 12.9 33.7 93.5 293.5 50.26

6 Intercooled (IC) cycle

The configuration of the engine is shown in Figure 8. The DP performance specifications


of the engine follow those shown on Table 2, and they are similar to those given for the
simple cycle. The results from DP calculations can be referenced in Figure 9. As first
approach, in Figure 8, the RC,total is splitted.
The fact that the cycle in question incorporates IC compression implies that the GT
contains two compression stages. For a given overall pressure ratio, if the Low Pressure
Compressor (LPC) operates at a high-pressure ratio and the High Pressure Compressor
(HPC), therefore, at a relatively low pressure ratio, the exit temperature of the HPC will
be relatively low, which means that to achieve a given ΤΕΤ a lot of heat must be supplied
in the combustor, resulting in high fuel consumption. The opposite is true for a small
pressure rise in the LP section and a high pressure ratio in the HPC, namely this will
result in lower fuel consumption. The latter configuration can therefore be expected to be
the more favourable one.
A range of overall pressure ratios was investigated with equal pressure ratio splits.
This should result (Saravanamutoo et al., 2001; Pilidis, 2006), in the highest efficiency
attainable with that particular overall pressure ratio. Bases on this, pressure ratios ranging
666 A.L. Polyzakisz et al.

from 42 to 90 were selected for more detailed examination, i.e., with various pressure
ratio splits. Each data point constitutes a particular pressure ratio split between the LP
and HP compressor. Several runs showed that an overall pressure ratio of 12.6 gives the
best performance in terms of SFC. The pressure ratio split which gives minimum SFC is
found to be RLPC = 1.4 and consequently RHPC = 9. This is very different from the
thermodynamically ideal situation of RLPC = RHPC = 3.55. The explanation for this lies in
the fact that the ideal case assumes perfect components. As soon as component
efficiencies are introduced, the efficiency rises as the LP ratio drops.

Figure 8 Intercooled cycle 2-shaft GT layout. ‘Brick’ and ‘station vector’ model

Figure 9 Intercooled cycle Design Point performance

As already mentioned, the GT selection for CC duty differs greatly from the optimum SC
machine presented above. The EGT of the latter is 573 Κ. This is far too low for efficient
heat recovery in a steam bottoming cycle. Suffice it to say that the overall pressure ratio
has to be much lower so as to end up with a much higher EGT. GTs with pressure ratios
between 8 and 24 were simulated in TurboMatch, which brought the EGT up to values of
773–873 Κ.
The choice of configuration will influence the DP performance inasmuch as that the
number of components and therefore the cumulative losses induced by all the
components, as well as the complexity and inherent cost, varies from one option to the
other. As far as OD performance is concerned, it can be said that a multi-shaft machine in
general is more flexible than a single-shaft machine, because its gas generator is allowed
to operate at a different rotational speed from that required for the PT. When the machine
is used for heavy-duty base load power generation, the PT speed is fixed by the required
frequency of the electric output from the generator – a gearbox is not desirable, for this it
would introduce additional cost and losses and so negatively influence the performance.
An engine with a PT obtains its power variation by varying the gas generator speed.
This means that torque (i.e., load output) changes are inherently slow and dependent on
rotor inertia, surge margin and temperature limits. Οn the other hand, single-shaft engines
Energy analysis of Brayton combined cycles 667

can handle rapid load changes once synchronised, as the rotational speed will always
remain unchanged and, therefore, there is nο rotor inertia to overcome. Furthermore,
it has a higher resistance to speed change in case of an unexpected change in electric
load because of its high rotor inertia. Summarising, a single-shaft machine is more
speed-stable and therefore easier to regulate, while at part load a two-shaft engine gives
better performance.
As in the case of the simple cycle, OD performance simulation was carried out.
The simulation refers to the variation of ambient temperature, altitude and TET
(Figures 10–13).

Figure 10 Power vs. TET varying Tamb (IC)

Figure 11 Power vs. TET varying altitude (IC)

Figure 12 Efficiency vs. TET varying Tamb (IC)


668 A.L. Polyzakisz et al.

Figure 13 Efficiency vs. TET varying altitude (IC)

Similar to the simple cycle, the optimum DP of the GT alone is very different
from the one eventually selected for combined cycle duty (Table 4) (Polyzakis et al.,
2007). There is an optimum solution where neither the GT thermal efficiency
nor the steam plant power output are the best achievable, but which yields optimum
CC thermal efficiency. The reason for this is the fact that when the efficiency
of a GT drops, its EGT increases. The additional heat allows more steam to be generated
in the HRSG, thereby increasing the steam turbine power output. It must be also
noted that the selection of the more suitable engine is taking into account the
possible aerodynamic problems which might occur in the GT compressor when the
RC overall is divided in an extremely unequal way between the two compressors.
From Table 4, it can be seen that the optimum pressure ratio (RC) for the CC occurs in the
region 12.6.

Table 4 Selection of optimum cycle with intercooling GT (TET = 1,561 K)

ηGT Exh.flow EGT XACT PSTEAM CCPP ηCCPP


RC,TOTAL RC,LPC × RC,HPC (%) (kg/s) (K) (%) (MW) (MW) (%)
1.25 × 8 31.53 687.7 849.1 11.8 110.4 310.4 48.9
1.5 × 6.667 31.75 664.6 845.5 11.7 105.9 305.9 48.6
1.725 × 5.8 31.76 651.6 842.9 11.7 103.3 303.3 48.2
2×5 31.64 641.4 840.3 11.7 101.2 301.2 47.6
10
2.25 × 4.444 31.46 635.7 838.2 11.7 99.8 299.8 47.2
2.5 × 4 31.23 632.4 836.4 11.7 104.3 304.3 47.5
2.75 × 3.636 30.97 631.0 834.9 11.8 103.6 303.6 47.0
3 × 3.333 30.69 630.7 833.5 11.9 103.1 303.1 46.5
1.26 × 10 33.59 669.8 813.5 12.8 103.1 303.1 50.9
1.4 × 9 33.79 654.2 811.4 12.9 100.0 300.0 50.7
1.575 × 8 33.91 639.7 809.0 13.0 97.1 297.1 50.4
12.6 2.1 × 6 33.80 615.3 803.8 13.2 91.8 291.8 49.3
2.52 × 5 33.47 606.5 800.6 13.4 89.6 289.6 48.5
3.15 × 4 32.81 602.1 797.0 13.6 88.0 288.0 47.2
3.55 × 3.55 32.35 605.5 795.1 13.7 87.5 287.5 46.5
Energy analysis of Brayton combined cycles 669

Table 4 Selection of optimum cycle with intercooling GT (TET = 1,561 K) (continued)

Exh.flow EGT XACT PSTEAM CCPP ηCCPP


RC,TOTAL RC,LPC × RC,HPC ηGT (%) (kg/s) (K) (%) (MW) (MW) (%)
1.25 × 12.8 35.40 664.9 778.7 14.5 91.5 291.5 51.6
1.6 × 10 35.87 627.4 773.7 14.8 84.9 284.9 51.1
2×8 35.89 604.5 769.5 15.0 80.6 280.6 50.4
16 2.4 × 6.667 35.67 591.9 766.3 15.1 78.1 278.1 49.6
3.2 × 5 34.90 581.8 761.5 15.4 75.5 275.5 48.1
3.6 × 4.444 34.46 580.7 759.6 15.5 74.8 274.8 47.4
4×4 34.00 581.2 758.0 15.6 74.5 274.5 46.7

7 Reheated (RH) cycle

Α substantial increase in specific work output can be obtained by splitting the expansion
and reheating the gas between the high pressure and low pressure turbines. At the same
time, the thermal efficiency of the GT will be reduced in the region of relatively
low pressure ratios; on the other hand, the heat ‘rejected’ to the HRSG will be increased,
due to the increased EGT. The configuration of the engine is shown in Figure 14. As a
first approach, the RC,total is spitted. The specifications of the engine follow those shown
on Table 2. This makes the specifications of the engine the same as those of the state
of the art and it gives the opportunity of a comparison between the results of the design
project with the manufacturer’s specifications. The results from DP calculations can be
seen in Figure 15. From the point of view of aerodynamical and mechanical design of the
compressors, it is better tο split the overall Rc as equally as possible 7 × 7 rather than
35 × 1.4. This is because the LPC will experience more severe surge problems and the
LPC is more sensible in such sort of problems than HPC. On the other hand, eνen with
the split 7 × 7, variable guide vanes should be used for both compressors (due tο Rc > 5
in both of them), but the surge problems will not be so severe for the LPC as previously.
But, from the DP of view, the higher the LPC pressure ratio, the better the efficiency
(Polyzakis, 1995; Mattingly, 1999; Arrieta and Silva Lora, 2005; Walsh and Fletcher,
1999; Kim et al., 1994).
Having decided οn the DP, the next step is to investigate the behaviour of the engine,
how it would perform over a wide range of ambient temperatures, altitude and different
loads (TETs) (Figures 16–19). The selection of the appropriate engine is based on the
comparison of the performance of different GTs with a reheat engine. Based on the GT
output simulation data, steam SC calculations have been carried out.

Figure 14 Reheated cycle 2-shaft, 2-spool GT layout. “Brick” and “station vector” model
670 A.L. Polyzakisz et al.

Figure 15 Reheated cycle Design Point performance

Figure 16 Power vs. TET varying Tamb (RH)

Figure 17 Power vs. TET varying altitude (RH)

Figure 18 Efficiency vs. TET varying Tamb (RH)


Energy analysis of Brayton combined cycles 671

Figure 19 Efficiency vs. TET varying altitude (RH)

Ιn Table 5, the overall results of the GT cycle, steam cycle (steam flow, steam cycle
efficiency, steam wetness, steam cycle power output) and the overall CCPP power and
efficiency are included. It can be seen that the optimum pressure ratio (RC) for the CC
occurs in the region 49 (Polyzakis et al., 2007).

Table 5 Selection of optimum GT with reheat (TET = 1,561 K)

RC Exh.flow EGT St.flow XACT ηSTEAM PSTEAM CCPP ηCCPP


(overall) RC,LPC × RC,HPC ηGT (%) (kg/s) (K) (kg/s) (%) (%) (MW) (MW) (%)
30 15 × 2 35.87 551.9 869.5 80.47 9.9 36.1 99.3 299.3 53.7
42 11.05 × 3.8 34.86 527.8 915.5 82.17 7.8 36.6 107.2 307.2 53.5
42 14 × 3 35.70 540.0 885.0 80.77 9.2 36.3 101.3 301.3 53.8
11.66 × 4.2 34.56 534.9 917.6 84.17 7.7 36.7 109.2 309.2 53.4
49 12.25 × 4 34.80 536.3 911.1 83.54 8.0 36.6 107.7 307.7 53.5
22.27 × 2.2 36.68 582.7 823.6 78.22 12.0 35.5 91.9 291.9 53.5
54 13.5 × 4 34.81 544.3 902.0 83.62 8.4 36.5 106.7 306.7 53.4
57 14.25 × 4 34.76 550.3 897.6 83.97 8.6 36.4 106.7 306.7 53.3
64 16 × 4 34.65 563.4 887.6 84.62 9.1 36.3 106.4 306.4 53.1
64 21.33 × 3 35.84 582.9 841.8 60.97 11.2 35.7 97.0 297.0 53.2

8 Intercooled and Reheated (IC/RH) cycle

There are several possible engine configurations corresponding to the IC/RH cycle.
The one investigated here is shown in Figure 20. The DP performance specifications of
the engine follow those shown in Table 2, and they are similar to those given for the
simple cycle. The results from DP calculations can be referenced in Figure 21.
Once the optimum IC/RH GT was selected, the analysis of the OD performance
behaviour of the engine was carried out. Α wide range of operational conditions (ambient
temperature, altitude, TET) were investigated (Figures 22–25).
672 A.L. Polyzakisz et al.

Figure 20 Intercooled reheated cycle 3-shaft, 2-spool GT layout. ‘Brick’ and ‘station vector’
model

Figure 21 Intercooled and reheated cycle DP performance

Figure 22 Power vs. TET varying Tamb (IC/RH)

Figure 23 Power vs. TET varying altitude (IC/RH)


Energy analysis of Brayton combined cycles 673

Figure 24 Efficiency vs. TET varying Tamb (IC/RH)

Figure 25 Efficiency vs. TET varying altit (IC/RH)

Nonetheless, by assuming for the sake of the comparison that overall pressure ratio
was split into equal proportions between the two compressors, some trends could be
discovered.
Firstly, as overall pressure ratio increased, the GT efficiency increased and
simultaneously the EGT decreased. The two effects have contradictory influence on the
CC efficiency and tend to balance each other, producing a maximum, at an intermediate
value of pressure ratio.
Secondly, as overall pressure ratio increase, SP increases and so did the CC
efficiency. This is valid until certain value of pressure ratio, beyond which the EGT
reduction starts counteracting and efficiency, begins to decrease. At the same time, the
steam power output continuously decreases as pressure ratio increase, due to the
diminishing of EGT. On the other hand, if the overall pressure ratio is kept constant at
values lower than 50, the gas SP seems to play a role in the CC efficiency. This increased
as gas SP went up (Velez, 1995). The reason for this is related to the simultaneous
increasing of the EGT, which means that more heat was still available to be transformed
in useful power into the steam turbine.
For high values of overall pressure ratio things were rather different. In this case a
relationship between the LP compressor pressure ratio and CC efficiency was readily
visible. The lower the LP pressure ratio, the higher the CC efficiency. This tendency was
valid independent of the value of overall pressure ratio. Because a reheating system was
674 A.L. Polyzakisz et al.

working at the inlet of the LP turbine, the EGT could only be influenced by the LP
compressor pressure ratio. This means that as the LP pressure ratio decreased, the EGT
increased and, consequently, so did the CC efficiency. The simultaneous reduction in GT
efficiency could not mitigate the stronger effect of EGT.
The above outlined observations confirmed the initial assumption that the best
GT, from the stand point of fuel consumption is not necessarily the engine that best
matches a steam cycle in a CCPP. The final result of this analysis was the selection of an
IC/RH GT having a pressure ratio of 60 (4.5 × 13.33) and fired at 1,508 Κ, (Table 6)
(Polyzakis et al., 2007).

Table 6 Selection of optimum cycle with intercooling and reheat GT (TET = 1,508 K)

SPGT Exh.flow St.flow XACT ηSTΕΑΜ PSTEAM CCPP ηCCPP


RC,TOTAL RC,LPC × RC,HPC ηGT (%) (J/kg) (kg/s) EGT (K) (kg/s) (%) (%) (MW) (MW) (%)
2 × 25 28.85 400,800 450.3 1,077 1,077 13.1 35.59 12.3 320.3 46.1
5 × 10 36.57 492,600 520.2 871 871 11.3 35.58 92.7 292.7 50.52
50 6.5 × 7.69 36.68 490,200 421.8 831 831 11.2 35.51 67.9 267.9 49.12
7.07 × 7.07 36.50 486,600 424.7 820 820 11.2 35.49 66.3 266.3 48.61
8 × 6.25 36.18 480,800 427.8 804 804 11.0 35.09 63.2 263.2 47.62
3 × 20 34.96 463,000 449.9 978 449.9 12.5 37.71 106.8 306.8 53.63
5 × 12 36.92 500,000 416.7 871 416.7 11 36.93 77.0 277.0 51.14
60 6 × 10 37.22 501,300 415.6 840 415.6 10.9 35.75 68.9 268.9 50.04
7.74 × 7.74 37.04 495,000 420.6 802 420.6 10.9 35.07 61.8 261.8 48.48
10 × 6 36.06 478,500 422.7 771 422.7 10.6 33.96 54.9 254.9 45.96
4 × 17.5 35.40 490,200 425.1 919 425.1 12 36.94 87.6 287.6 50.91
7 × 10 37.60 506,300 411.4 815 411.4 11.2 35.61 63.6 263.6 49.56
8.36 × 8.36 37.33 500,000 416.5 789 416.5 10.9 34.92 58.7 258.7 48.28
70
10 × 7 36.79 490,200 424.1 765 424.1 10.8 33.62 53.5 253.5 46.63
11 × 6.36 36.27 481,900 431.9 754 431.9 10.8 33.49 52.4 252.4 45.78
17.5 × 4 32.76 430,100 483.2 706 483.2 10.5 31.29 46.4 246.4 40.35

9 Intercooled, Reheated and Regenerative (IC/RH/HX) cycle

There are several possible engine configurations corresponding to the IC/RH/HE cycle.
The one investigated here is shown in Figure 26. The DP performance specifications of
the engine follow those shown in Table 2, and they are similar to those given for the
simple cycle. The results from DP calculations can be referenced in Figure 27.
A detailed investigation, shown that the engine with the best efficiency at the
DP performance has the following specifications: TET = 1,561 K, Rc(overall) = 31,
(RC,LPC)/(RC,HPC) = 0.8, and performance: ηGT = 40.05%, SPGT = 447,487 J/kg, Exh. Flow
= 446.94 kg/s.
Once the optimum IC/RH/HE GT was selected, the analysis of the OD performance
behaviour of the engine was carried out. Α wide range of operational conditions (ambient
temperature, altitude, TET) were investigated (Figures 28 and 29).
Energy analysis of Brayton combined cycles 675

Figure 26 Intercooled reheated heat exchanger cycle 3-shaft, 2-spool GT layout. ‘Brick’
and ‘station vector’ model

Figure 27 Intercooled and reheated cycle Design Point performance

Figure 28 Power vs. TET varying Tamb (IC/RH/HE)

Figure 29 Power vs. TET varying altitude (IC/RH/HE)


676 A.L. Polyzakisz et al.

It must be stated that an engine such as this is not suitable for use in a practical combined
cycle since the heat exchanger lowers the temperature of the exhaust gases too much for
the steam cycle to be of sufficiently high thermal efficiency.
The other four engines were compared after being optimised for combined cycle
mode using a single pressure steam cycle as the yard stick. Several selection parameters
were used in making the choice.

10 Engine selection: multi criteria analysis

The final target of this work is the selection of the engine with the best performance from
the point of view of various technical and economical considerations. A number of
criteria were chosen to give an impression of the CC performance. The criteria that were
employed for this study were, in order of importance (Pilavachi, 2002; Kim et al., 1994;
Najjar, 1999):
• Combined cycle efficiency, indicative of long-term profitability of the plant.
• Cost, entailing the GT layout and complexity, and its exhaust flow, which determines
the size of the steam plant (Lazaro and Millan, 2006).
• GΤ efficiency. A CCPP will usually operate οn GT alone until the steam plant is
completed or during steam system downtime.
• Boiler pressure, which gives an indication of the steam plant complexity
and ST plant cost.
• TΕΤ, indicating the life and reliability/availability of the GT.
• OD performance.
Many other criteria can be mentioned such as EGT, exit mass flow, maintenance, sale
policy, etc. The final selection was made using a matrix method employing weight
factors.
Figures 30, 31 and Table 7 lists the main performance characteristics for the cycles
investigated. The next step is tο compare this engine with different cycle engines, the
specifications of which are similar.

Figure 30 Efficiency vs. TET varying Tamb (IC/RH/HE)


Energy analysis of Brayton combined cycles 677

Figure 31 Efficiency vs. TET varying altit (IC/RH/HE)

Table 7 Comparison of Gas Turbine cycles for CCPP

SC IC RH IC/RH IC/RH/HX
RC (overall) – 16 12.6 49 60 31
Exh.flow Kg/s 716.4 654.2 536 418.5 457.9
ηGT % 33.46 33.79 34.8 36.89 42.05
ηST % 34.5 35.6 36.6 36.93
Boiler pressure kPa 8,000 8,000 8,000 8,000
Cont pressure kPa 6 4 4.5 5
XACT % 12.8 12.9 8 11
CCPP Power MW 301.1 300.0 307.7 277
ηCCPP % 50.37 50.7 53.5 51.14

The use of a multi-objective optimisation made the selection of the Brayton cycle most
suitable for a combined cycle. The technique consists in choosing the most salient
parameters that influence performance of the alternatives examined. Each of these
characteristics is then evaluated on its own, for every alterative, and a score is assigned to
it on a common scale basis. The engine having at the end the highest score, included all
the criteria, is deemed to be the most feasible for the successive stages of the evaluation.
The relative importance of each criterion is represented by the adoption of an appropriate
weighting factor (Table 8).
The method is very sensitive to the weighting factors and to the score assigned to
each alternative for each criterion, therefore it must be considered as a first approach of
the selection. The final result is very dependent on the experience of the persons doing
the analysis. In particular the following procedure is adopted in this paper.
Table 8 shows the weight factors that were eventually assigned to each of these
criteria. Next, the cycles were ranked for each of these criteria. In the case of points 1 and
3 the cycle with the best performance was given 100 points and the others were scaled
according to their relative performance. The cost was evaluated by giving the cycle with
the simplest layout simple cycle 10 points, IC 8 points, RH 6 points, and IC/RH 4 points.
Then these numbers were multiplied by factors expressing the relative size of the steam
plant. This factor was calculated by dividing the smallest mass flow by each exhaust
mass flow. The OD performance was taken into account by giving 10, 9, 8 and 7 points to
the IC/RH, IC, RH and simple cycle GT, respectively.
678 A.L. Polyzakisz et al.

Table 8 Selection of optimum GT cycle for CCPP

ηCC Cost ηGT Boiler Pr. TET OD TOTAL


Rank 1 2 3 4 5 6
Weight factor 0.296 0.222 0.185 0.185 0.074 0.038 1.0
SC 94 115 93 90 94 79 2
27.88 25.54 17.15 16.74 6.92 3.01 97.24
IC 95 76 93 113 94 102 3
28.06 16.77 17.24 20.93 6.92 3.87 93.79
RC 100 123 96 90 94 90 1
29.60 27.29 17.83 16.74 6.92 3.44 101.82
IC/RH 96 71 102 91 104 113 4
28.30 15.80 18.91 16.75 7.69 4.30 91.75

Finally, these rankings were all scaled to give the same sum of all individual rankings
for each column, i.e. 385. They were then multiplied by the appropriate weight factor and
added up for each cycle configuration. The cycle with the most points is the most
favourable according to this analysis. It is clear that the RH GT cycle comes out as the
most suitable cycle for CCPP.
Ιn the case of the GT with reheat, the problem lies in the part load. The problem is
located mainly to the LPC, which is forced to go towards the surge line. There are two
solutions to that problem. The first is to put bleed ducts at the end of the LPC extracting
the excess amount of air. This solution is simple and it has low initial cost but reduces the
thermal efficiency of the GT, because power is consumed to compress at LPC air
and some amount of this compressed air simply goes away. The second, which is also
relatively simple and more appropriate solution, is to install at the inlet of the LPC
variable guide varies. The requirement of decreased non-dimensional mass flow
and higher surge margin is achieved without waste of energy.
The RH GT cycle came out to be the most suitable cycle for CC application.

11 Conclusions

This paper contains information regarding the optimisation of combined cycle power
plants. The aspects investigated are GT performance, steam turbine performance and then
the effects of both combined. The software developed has the ability to select
the optimum GT configuration for detailed study. Aspects which can be varied are
pressure ratio, turbine entry temperature, the effects of reheating an engine, the steam
boiler pressure, the condenser pressure and properties within the heat recovery boiler
such as exhaust gas approach temperature, evaporator pinch point, evaporator approach
point and feedwater temperature.
As the name implies, combined cycles are the combination of two power cycles: a GT
(Brayton) cycle and a steam (Rankine) cycle. The two cycles have quite different
characteristics: the gas cycle is a high temperature cycle with TET’s from 1,373 K to
1,573 K and heat rejection around 773 K; the steam cycle is a low temperature cycle with
a maximum temperature of about 823 K and heat rejection at ambient temperature.
Energy analysis of Brayton combined cycles 679

It was seen that the best combined cycle plant is far from being composed of the best
GT coupled with the best steam cycle. GTs will certainly play a major role in future
power generation and several well-justified concepts have been developed or are the
subject of major feasibility studies. However, the combined cycle is now well established
and offers superior performance to any of the competing systems which are likely to be
available in the medium term for large scale power generation applications. This work
contains an optimisation analysis of four potential GT cycles, namely, SC, IC, RH,
IC/RH, IC/RH/HX cycle.
The DP and OD point performance of the four cycles alone or with the cooperation of
the steam cycle has been investigated. The three options, SC, IC and RH perform very
similarly OD. The thermal efficiencies and SP outputs are not the optimum values they
would be for GT cycle operation. Furthermore, the IC/RH engine was run differently
from the IC one. When ambient temperature changes, it is prudent to assume that the
intercooler exit temperature also changes. Cooling is usually achieved with water or air
drawn from the environment, which will vary in temperature. This aspect was included in
the IC GT OD simulation, but not in the other two machines incorporating intercooling.
This explains the remarkably better performance of the latter two. At elevated ambient
temperature, the cooling water temperature would increase simultaneously.
Consequently, the HP compressor inlet temperature would be higher than in the DP
situation. This adversely affects the GT performance. It must be stated that an engine
such as IC/RH/HX is not suitable for use in a practical combined cycle since the heat
exchanger lowers the temperature of the exhaust gases too much for the steam cycle
to be of sufficiently high thermal efficiency. For that reason, the combined cycle
calculations in this report are of academic interest only. However, it must be noted that
(IC/RH/HX) GT is more efficient than any of the others. This will be the case for the DP
as well as in OD.
The other four engines were compared after being optimised for combined cycle
mode using a single pressure steam cycle as the yard stick. The optimum GT cycle to
operate in a combined cycle power plant came out to be the reheated cycle. This is due to
high EGT = 911.1 K and steam flow = 83.54 Kg/s the steam cycle efficiency is 36.6 and
the overall efficiency is 53.5. It is worth mentioning that a fraction of wetness is 8.0
much lower than the acceptable limit of 13, which simply means longer life for the
installation. That parameter and also the increased surge problems to the LPC, made the
14 × 3 GT reheated option not so attractive as the one that was selected. The total power
output is 307.7 MW, exceeding the need of 300 MW and that gives the opportunity to
‘scale’ downwards since the size of the GT will decrease costs.
The important thing learnt here was that the best combined cycle is not made up of
the best GT and the best steam cycle. It is not a straightforward matter to pick the
optimum combined cycle since so many parameters have to be taken into account.
A combined cycle plant is dependent on high specific work output and high EGT for use
in the steam cycle. High specific work output and high EGT contradict each other, hence
an optimum can be achieved. The less efficient GT is compensated by having a high EGT
thus producing a more efficient steam cycle. It is the combination of gas and steam
turbine efficiency that needs to be considered when assessing the overall efficiency of
combined cycle plant.
CCPP provide operators with less risk. The time to complete a CCPP plant is in
the region of 3–4 years. Coal fired plants will more like 6–8 years. Estimates required for
CCPP cover a short time span, so have a great probability of being accurate. It is also
680 A.L. Polyzakisz et al.

possible for the operator to use the GTs in open cycle to produce power, while the steam
cycle is being built. This enables revenue to be obtained from the project between the
first and second year. Capital outlay is reduced and cost per kilowatt kept to a minimum.
As extensive developments improve the reliability of GT availability figures
for CCPP are approaching 98–99% minimising downtime to maximise operators’ profit.
With reduced manning levels operating costs can also be reduced, if fuel costs were
assumed to be similar. A 600 MW coal fired power station would have manning levels in
the region of between 200 and 300 staff, while a comparable CCPP plant would require
between 40 and 60 staff. As the statements outlined above emphasise CCPP with regard
to power generation it is of great importance, and while natural gas prices remain low it
will continue to be.

References
Alabdoadaim, M.A. (2006) ‘Performance analysis of combined Brayton and inverse Brayton
cycles and developed configurations’, Applied Thermal Engineering, Vol. 113, No. 26,
pp.1448–1454.
Arrieta, F.R. and Silva Lora, E.E. (2005) ‘Influence of ambient temperature on combined-cycle
power-plant performance’, Applied Energy, Vol. 134, No. 80, pp.261–272.
Asea Brown Boveri (2004) Power Generation Review and Specifications Brochures,
ABB Publication, Power Production Group, UK.
Bolland, O.Α. (1995) ‘Comparative eva1uation of advanced combined cyc1e alternatives’,
Journal of Engineering for Gαs Turbines αnd Power, Vol. 113, No. 35, pp.190–197.
Fielden, C. (1995) Industrial GT for Combined Power Generation–Optimisation Issues,
MSc Thesis, Cranfield University, UK.
GE Energy (2004) F System, GE Publication, GE Energy, UK.
GE Power Systems (2004) GE Gas Turbines Specification Brochures, GE Publication, GE Power
Systems, UK.
Hans Dieter Baehr (1999) Thermodynamics, 3rd ed., Springer-Verlag Heidelberg, Berlin,
New York.
Heppenstall, Τ. (1998) ‘Advanced gas turbine cycles for power generation: a critical review’,
Applied Thermal Engineering, No. 18, pp.837–846.
Khaliq, A. and Kaushik, S.C. (2004) ‘Second-law based thermodynamic analysis of Brayton/
Rankine combined power cycle with reheat’, Journal of Applied Energy, Vol. 110, No. 78,
pp.179–197.
Kim, T.S. and Hwang, S.H. (2006) ‘Part load performance analysis of recuperated gas turbines
considering engine configuration and operation strategy’, Journal of Energy, Vol. 121, No. 31,
pp.260–277.
Kim, T.S. and Hwang, S.H. (2007) ‘Design and off design characteristics of the alternative
recuperated GT cycle with divided turbine expansion’, Transactions of ASME, Vol. 129,
April, pp.428–435.
Kim, T.S., Oh, C.H. and Ro, S.T. (1994) ‘Comparative analysis of the οff design performance for
gas turbine cogeneration systems’, Transactions of ASME, Vol. 14, Νο. 2, pp.153–163.
Kolanowski, B.F. (2003) Small-Scale Cogeneration Handbook, The Fairmont Press, Inc., UK.
Lazaro, E.C. and Millan, A.R. (2006) ‘Analysis of cogeneration in present energy framework’,
Fuel Processing Technology, Vol. 93, No. 87, pp.163–168.
Energy analysis of Brayton combined cycles 681

Leenhouts, F. (1995) Industrial GT for Combined Power Generation and Other Advanced Power
Generation Systems, MSc Thesis, Cranfield University, UK.
Mahi, Ρ., Gibson, C. and Morton, Α. (1994) ‘CCGTs: have they improved the environmenta1
situation in the UK’, Proceedings: IMechE Seminar, Commissioning and Operation of
Combined Cycle Plant, November, pp.73–86.
Matta, R.K. (2000) Power Systems for the 21st Century – ‘H’ GT Combined-Cycles, GE Power
Systems, GER-3935B (10/00), USA.
Mattingly, J.D. (1999) Elements of Gas Turbine Performance, McGraw-Hill editions International,
UK.
Najjar Y.S.H. (1999) ‘Comparison of performance of the IGSC with the CC’, Journal of Applied
Thermal Engineering, Vol. 3, No. 19, pp.75–87.
Najjar Y.S.H. (2001) ‘Efficient use of energy by utilizing gas turbine combined systems’,
Applied Thermal Engineering, No. 21, pp.407–438.
Palrner, J.R. (1995) The TurboMatch Scheme for Gαs Turbine Performance Calculations – User’s
Guide, Cranfield University, UK.
Pasha, Α. and Jolly, S. (1995) ‘Combined cycle heat recovery steam generators – optimum
capabilities and selection criteria heat recovery systems & CHP’, Transactions of ASME,
Vol. 2, No. 15, pp.147–154.
Pilavachi, P.A. (2002) ‘Mini and micro-GTs for CHP’, Journal of Applied Thermal Engineering,
No. 22, pp.2003–2014.
Pilidis, P. (2006) GT Performance Lecture Notes, Cranfield University, Cranfield.
Polyzakis A.L. (1995) Performance of Industrial GT for CCPP, MSc Thesis, Cranfield University,
UK.
Polyzakis, A. (2007) Technoeconomic Evaluation of Trigeneraton Plant: Gas Turbine
Performance, Absorption Cooling and District Heating, PhD Thesis, Cranfield University,
UK.
Polyzakis, A., Koroneos, C. and Xydis, G. (2007) ‘Optimum gas turbine for combine cycle power
plant’, International Journal of Energy Conversion and Management, Accepted.
Polyzakis, A.L. (2006) GT Performance and Applications, Kozani’s Institute of Technology
publications, Kozani.
Poulikas, A. (2004) ‘An overview and future sustainable gas turbine technologies’, Renewable and
Sustainable Energy Reviews, No. 9, pp.409–443.
Rice, Ι.G. (1986) ‘Thermodynamic evaluation of gas turbine cogeneration cycles: Part Ι – heat
balance method analysis’, Journal of Engineering for Gas Turbines and Power, No. 109,
pp.1–15.
Rufli, Ρ. (1987) ‘Α systematic ana1ysis of combined gas/steam cyc1e’, ASME GT, Vol. 1,
pp.135–146.
Saravanamutoo, H., Rogers, C. and Cohen, H. (2001) GT Theory, 5th ed., Logmann Scientific
& Technical, New York.
Snin, J.Y. and Jeon, Y.J. (2002) ‘Analysis of the dynamic characteristics of a CCPP’, Energy,
Vol. 27, pp.1085–1098.
Tinsley, D.G. (1995) Performance of Industrial GT for CCPP, MSc Thesis, Cranfield University,
UK.
Velez, L. (1995) Industrial GT for Combined Cycle Power Plant, MSc Thesis, Cranfield
University, UK.
Walsh, P. and Fletcher, P. (1999) GT Performance, 2nd ed., Blackwell Science Ltd., Oxford.
682 A.L. Polyzakisz et al.

Nomenclature
CC Combined Cycle (–)
CCPP Combined Cycle Power Plant (–)
CP Specific heat (kJ/(kg.K))
CW Compressor Work (MW)
DP Design Point (–)
D Difference (–)
S Entropy (J/mol/K)
EGT Exhaust Gas Temperature (T)
FCV Fuel Calorific Value ()
GT Gas Turbine (–)
HI Heat input (MW)
HRSG Heat Recovery Steam Generator (–)
HP High Pressure (–)
IC Intercooled Cycle (–)
IC/RH Intercooled and Reheated Cycle (–)
IC/RH/HX Intercooled, Reheated and Regenerative Cycle (–)
LP Low Pressure (–)
M Mach number (–)
m Mass flow (kg/s)
MIXEES Mixer (–)
NG Natural Gas (–)
NOZCON Convergent Nozzle (–)
OD Off Design (–)
Rc Overall Pressure Ratio (–)
PREMASS Air Mass Splitter (–)
PT Power Turbine (–)
RH Reheated Cycle (–)
SC Simple Cycle (–)
SFC Specific Fuel Consumption (g/MJ)
SP Specific Power (kJ/kg)
SW Specific work (MJ/kg)
Tstack Stack Gas Temperature (K)
TET Turbine Entry Temperature (K)
TW Turbine Work (MW)
UW Useful Work (MW)
XACT Wetness (–)
η Efficiency (–)
Energy analysis of Brayton combined cycles 683

Subscripts–Superscripts
amb Ambient
c Compressor
cc Combustion Chamber
CC Combined Cycle
ch Choked
cold Cold section of the engine (i.e. intake, compressor)
in Inlet
eh
f Fuel
hot Hot section of the engine (i.e., burner, turbine, exhaust)
is Isentropic
loss Losses
m Mean
o Total or absolute values
od Off design
out Outlet
t Turbine
th Thermal
Greek Ambient
γ Gamma (–)
η Efficiency (–)

You might also like