You are on page 1of 88

THE CONSTRAINT ANALYSIS

OF A 250 kW MICROTURBINE
GENERATOR

By: Avnish Garg


Final Year
B.Eng. in Electromechanical Engineering
Aston University
Birmingham, U.K.

Report prepared for:


Dr. J. Hill
School of Engineering and Applied Science
Aston University
Birmingham, U.K.

1
ACKNOWLEDGMENTS

I am indebted to my parents, my family for their trust and support. I am highly


grateful to my supervisor, Dr. J. Hill for his keen interest in my work and his generous
support throughout the duration of this project. His encouragement and motivation
spurred me onto greater heights. His valuable comments, suggestions, criticisms
inspired me to give off my best and improve the quality of my project at every step.

2
SYNOPSIS

This project involves the development of a constraint (technical and economic) analysis for a
250 kWe microturbine engine as means for electric power generation. Various design
parameters have been identified and quantified in order to determine the economic viability
and performance characteristics of a 250 kW microturbine system. Microturbines are small-
scale distributed generation systems based on a similar technology to large-scale gas turbine
power plants but they work at low pressure ratios.

The key technical differences between a microturbine and a gas turbine are that most
microturbines (usually) use centrifugal (or radial flow) turbomachinery instead of axial
turbomachinery and a recuperator (heat exchanger) instead of a regenerator. A parametric
cycle analysis for the ideal and actual unrecuperated (simple) and recuperated gas turbine cycle
(known as the Brayton Cycle) has been carried out in order to evaluate the influence of
recuperation on the performance and fuel consumption of the microturbine. Effects on various
design and performance parameters have been analyzed while keeping the other input
parameters constant. Effects of different component efficiencies, pressure ratios and turbine
inlet temperatures on the overall efficiency, fuel costs and other design parameters have been
discussed. An effort has been made to determine optimum operating conditions for a 250 kW
microturbine system.

An attempt has been made to define the range of configurations and key design parameters that
are able to fulfil the requirements for a 250 kWe microturbine engine, hence defining the
design space. Key technical limitations of microturbines have been addressed and solutions for
these problems have been discussed. Current applications of microturbine systems have been
studied and possible future applications have been proposed.

3
CONTENTS

Synopsis
1 Introduction 5
2 Components of a Microturbine 6
2.1 Compressor 7
2.1.1 Centrifugal (or Radial Flow) Compressor 9
2.1.2 Axial Flow Compressor 11
2.2 Radial and Axial Flow Turbines 12
2.3 Combustion Chamber 14
2.4 Heat Exchanger 15
2.4.1 Regenerator 16
2.4.2 Recuperator 16
2.5 Other Components 18
3 Cycle Analysis 19
3.1 Ideal Simple Brayton Cycle 19
3.2 Ideal Brayton Cycle with Heat Exchange 28
3.3 Actual Simple Brayton Cycle 31
3.4 Actual Brayton Cycle with Heat Exchange 39
3.5 Electric Power Output and Shaft Power Required 43
3.6 Fuel Consumption and Costs 44
4 Conclusion 48
5 Nomenclature 50
6 References 52
7 Figures 54
8 Graphs 56

4
1. INTRODUCTION

In the recent decades, many factors have contributed to the development and usage of
distribution generation technologies, such as microturbines, as a means to reduce load
on large power plants globally. Newfound awareness about pollution control
measures and emission-free power generation has made distributed generation a
necessity. Rising fossil fuel prices, depletion of fossil fuels and ever-increasing
electricity consumption have led to a situation of ‘energy crises’ and necessitated the
development of and research into alternative sources of energy such as bio-fuels and
refuse (both industrial and domestic). Recent energy legislations such as the Kyoto
Protocol which limit the levels of carbon dioxide and other atmospheric pollutants for
various sectors have spurred on this development. Microturbines are just one of these
means of distributed generation which are bound to fulfil these future energy needs of
the world.

A Microturbine Generator may be described as a stationary gas turbine power plant


which predominantly produces on-site electric power (prime-mover application)
which has a design electric output rating in the range of about 25 kWe to 500 kWe a
(Shane, 2002, p. 29) while conventional gas turbines are usually classified as those
which have outputs of greater than about 500 kWe. Salient features of microturbines
which make them appealing to customers include fuel flexibility, low toxic emissions,
compactness and modularity. One of the chief advantages of microturbine generators
is their simple design because of which they can be used for a large number of
applications by just making a few design modifications. Other possible applications of
microturbine include Combined Heat and Power (CHP), Hybrid Electric Vehicles
(HEV) and Heating, Ventilation & Air Cooling (HVAC) although development of
microturbines for these applications is currently in its initial stages.

a
There is no complete consensus about the classification of a gas turbine as a microturbine or a
conventional gas turbine among various cited sources. Hence, a broad range of values of electrical
power outputs from various models has been mentioned by the author.

5
2. COMPONENTS OF A MICROTURBINE

A 250 kWe microturbine has been discussed in this study due to the feasibility of
manufacture of its components. In microturbines with lower electric power outputs,
the machinability of the components reduces greatly and tendency of flaws increases,
thereby reducing the overall efficiency of the microturbine and increasing need of
maintenance and repair. Hence, a mid-range microturbine generator seems to be an
appropriate model for this study.

The key components of a typical microturbine are as follows (Stares and Mabbutt,
2002, pp. 67-72, Hamilton, 2003, p. 6-11):

a) Compressor (typically single-stage centrifugal type)


b) Turbine (typically single-stage radial flow type)
c) Combustion Chamber
d) Recuperator

Other components include the electric generator, bearings, power conditioning


systems, nozzles, and fuel filtration, metering and injection systems. These
components have not been analysed in this study. Additional components can be
added to the microturbine depending on the application of the system.

The microturbine generator (usually referred to as microturbine) works on principles


similar to larger gas turbines but is distinguishable from them with regard to a number
of technical aspects. Like large gas turbines, the thermodynamics of microturbines are
based on the Brayton cycle. The compressor and the turbine are mounted on a single
shaft along with the electric generator. Twin bearings support the shaft. The single
moving part of the single shaft design reduces the need for maintenance and enhances
overall reliability and durability.

The compressor compresses incoming atmospheric air. This air is then preheated in
the recuperator (only used in cycles with heat exchange) using heat from the exhaust
gases. The heated air from the recuperator is mixed with fuel in the combustion

6
chamber which increases the temperature, velocity and volume of the gas. The turbine
extracts work from the hot exhaust gases, thereby generating shaft power. This shaft
power drives the compressor and rotates the electric generator (which converts this
mechanical power to electric power). The frequency of this electric power can be
modified as per the requirements using power electronics systems.

The recuperated cycle is typically a popular choice (when disregarding the cost of the
recuperator) for microturbines because of its ability to achieve high thermal
efficiencies at low compression pressure ratios and its suitability to radial flow
turbomachinery. Currently, most microturbines work at low turbine inlet temperatures
of about 900°C and compression pressure ratios of about 4:1 (Pullen et al., 2002, p.
87). At such low pressure ratios, the thermal efficiencies of unrecuperated
microturbines are relatively low. Hence, the use of the recuperator is as such essential
for viable power generation.

2.1 COMPRESSOR

Gorla and Khan (2003, p. 6) describe a Compressor as a turbomachine which


performs work on a compressible working fluid (usually air, in the case of gas
turbines) by reducing its volume and then allowing it to gradually expand. This
process imparts kinetic energy to the fluid in the impeller and increases its internal
temperature and pressure energy in the diffuser. The kinetic energy imparted to the
working fluid forces the air through the nozzle to make the fluid suitable for efficient
combustion of the fuel.

Though many types of compressors are in use nowadays, the two major types of
compressors used in gas turbines (and microturbines) are the Centrifugal (or Radial
Flow) Compressor and the Axial Flow Compressor. According to Pullen et al. (2002,
p. 90), centrifugal compressors and mixed-flow compressors (hybrid of centrifugal
and axial flow compressors) are popular choices for compressors in microturbine
generators “on grounds of efficiency and development costs” at low temperature and
pressure conditions.

7
The ratio of the pressure at compressor exit to the pressure at compressor inlet
(usually atmospheric pressure in the case of gas turbines) is known as the
compression pressure ratio (Π) and plays an important role in the performance of the
microturbine system.

Centrifugal compressors are better suited for microturbines than axial ones due to the
following reasons (Energy Nexus Group, 2002, p. 3; Håll, 2002a, pp. 1-2;
Saravanamuttoo et al., 2001, p. 182):

a) Since microturbines need to use small turbomachinery, intricate designing, which


is an essential feature of axial compressors, is not feasible. Centrifugal
compressors are usually relatively more efficient at low pressure ratios (at which
microturbines work) due to their uncomplicated design, robustness and
insensitivity to flaws.

b) Centrifugal compressors can achieve higher pressure ratios per stage. This reduces
the number of stages and the size of the compressor significantly. Since size of
components is a critical factor for microturbines, centrifugal compressors are a
popular choice. Hence, centrifugal compressors in microturbines are usually
single stage models.

c) Centrifugal compressors can handle small volumetric flows with reasonably high
component efficiency. In the case of microturbines, blade height would be too
small to be practical for axial flow compressors.

d) In the size range of microturbines, centrifugal compressors offer minimum surface


and end wall losses thereby providing optimum efficiency.

Currently, the technology for centrifugal compressors is less developed than axial
compressors as the former had not found many applications before the advent of
microturbines. The complex design of the diffuser in centrifugal compressors reduces
their ease of manufacture. Another problem faced by centrifugal compressors is their
susceptibility to formation of deposits of particulate matter as this leads to corrosion
and reduction in durability of the compressor.

8
According to Stares and Mabbutt (2002, p. 68), compressor materials for
microturbines vary from cast iron and aluminium to stainless steel. Other robust
materials such as titanium and nickel alloys (such as Inconels, Monels and Nimonics)
can be used to provide better durability to the compressor.

2.1.1 CENTRIFUGAL (OR RADIAL FLOW) COMPRESSOR

According to Saravanamuttoo et al. (2001, p. 152-4) and Gorla and Khan (2003, p.
143-4), a centrifugal compressor basically consists of a rotating impeller (within a
stationary casing), which imparts a high velocity (with both radial and tangential
components) to the fluid, and a stationary diffuser (“a series of fixed diverging
passages”) which decelerates the air with a consequent increase in static pressure
(diffusion).

In a centrifugal compressor, air enters the impeller eye and is spun at a high velocity
by the vanes on the disc of the impeller, (thereby imparting a “centripetal
acceleration” and increased static pressure to the compressed air) and leaves and
enters the diffuser via the tip of the impeller. Usually, centrifugal compressors are
designed such that about half of the rise in static pressure of the air occurs in the
impeller and the remainder takes place in the diffuser.

According to Gorla and Khan (2003, p. 150), the basic purpose of the diffuser is to
reduce the velocity of the air which leaves the compressor and enters the combustion
chamber, thereby ensuring “efficient combustion of the fuel” while maintaining high
levels of static pressure in the air. Since air has a natural tendency to break away from
the walls of the diverging passage and flow back in the direction of pressure gradient,
it loses some of the static pressure developed in the impeller and the diffuser due to
formation of ‘Eddy Currents’. The angle of divergence of the passages in the diffuser
should not exceed a certain design value to limit stagnation pressure loss (∆pb).

9
Saravanamuttoo et al. (2001, pp. 163-4) state that to carry out the diffusion process in
as short a length as possible and to control the flow of the working fluid effectively,
the fluid leaving the impeller is divided by fixed diffuser vanes. During part-load
operation of the gas turbine, the flow direction varies with mass flow rate ( m ) and
pressure ratio (Π), thereby reducing the part-load efficiency of the gas turbine.
Usually, only large gas turbines use variable-angle diffuser vanes to adjust the flow
direction effectively over a wider range of operating conditions as incorporation of
this type of vanes increases the size and complexity of the system.

After leaving the diffuser vanes, the working fluid may be delivered to a volute and
then to the combustion chamber (via a heat exchanger, if used). A number of small
industrial gas turbines avoid using the diffuser vanes and instead use the volute alone
in order to reduce the compressor size.

Actual compression work done on the air and rise in pressure is affected by the
number of diffuser vanes (in terms of ‘slip’ factor σ), impeller tip speed (U);
‘windage’, disc friction and casing losses (in terms of power input factor ψ) and
compressor inlet temperature T1 (Saravanamuttoo et al., 2001, pp. 155-7; Gorla and
Khan, 2003, pp. 148-150).

According to Gorla and Khan (2003, pp. 145-6), the impeller tends to undergo high
stress forces. Curved blades (forward-curved and backward-curved) tend to straighten
out due to these centrifugal forces and bending stresses are set up in the vanes. Radial
blades are free from bending stresses and are “somewhat easier to manufacture than
curved blades”. Increased mass flow rate does not change the pressure on radial
blades but forward-curved blades can achieve higher pressure ratios.

Centrifugal stresses developed in the impeller of the single-stage centrifugal


compressor (usually using light alloys such as aluminium) limit the impeller tip speed
U to about 460m/s and the pressure ratio to about 4:1. According to Bullin (2002, p.
2), microturbine generators operate at pressure ratios of typically 4:1. Hence, it may
be inferred that centrifugal compressors are well-suited for them.

10
Currently, higher impeller tip speeds can be used with high-strength alloys such as
titanium or stainless steel and pressure ratios of 8:1 can be achieved but unsuitability
for mass production and low efficiencies associated with single-stage compressors act
as a deterrent to their popularity and increase costs. Additional losses in static
pressure and efficiency occur in compressors due to stalling (breakaway of flow) of
stages, surging (complete breakdown of continuous steady flow) and choking (of
mass flow at various stages) (Saravanamuttoo et al., 2001, pp. 155-7; Pullen et al.,
2002, p. 90; Gorla and Khan, 2003, p. 153-55, 187).

2.1.2 AXIAL FLOW COMPRESSOR

According to Saravanamuttoo et al. (2001, p. 182), “the axial flow compressor


consists of a series of stages, each stage comprising of a row of rotor blades followed
by a row of stator blades. The working fluid is initially accelerated by the rotor blades,
and then decelerated in the stator blade passages wherein the kinetic energy
transferred in the rotor is converted to static pressure. The process is repeated in as
many stages as are necessary to yield the required overall pressure ratio.”

In essence, the components of the axial compressor are somewhat analogous to the
centrifugal compressor where the rotor blades replace the impeller and the stator
blades replace the diffuser. The flow of the working fluid is axial (in the direction of
the axis of the compressor) to the axis of rotation (of the blades) as compared to the
radial (centripetal) direction of fluid flow in the centrifugal compressor.

As the pressure increases in the direction of flow, the volume of working fluid
decreases. The height of the blades is decreased along the axis of the compressor to
keep the axial velocity of the fluid approximately constant for each stage. The fluid is
directed at the correct angle onto the first stage by a row of inlet guide vanes. As the
number of stages (N) in the axial compressor increases, the overall pressure ratio
which can be achieved by it increases (Saravanamuttoo et al., p. 187).

11
Saravanamuttoo et al. (2001, p. 188) state that a high stagnation temperature rise
reduces the number of stages for a given pressure overall pressure ratio. In order to
ensure this, the following features need to be incorporated in the axial compressor – (a)
high blade speed; (b) high axial velocity; and (c) high fluid deflection in rotor blades.
The blade speed is limited by blade stresses and adverse pressure gradient and
aerodynamic considerations combine to limit axial velocity and fluid deflection.

According to Gorla and Khan (2003, pp. 187-8) one of the important characteristics of
the axial flow compressor is its ability to achieve high pressure ratios at good
efficiency. Modern axial flow compressors can provide efficiencies of 86-90%. Since
large gas turbines work at high pressure ratios and are not affected significantly by the
size of turbomachinery, axial compressors are suitable for such turbines. Axial
compressors (especially multi-stage) are unsuitable for microturbines since
compactness of turbomachinery is a very important requirement.

2.2 RADIAL AND AXIAL FLOW TURBINES

Gorla and Khan (2003, p. 283) describe a turbine as a machine which extracts kinetic
energy from hot expanding gases leaving the combustion chamber, converting this
kinetic energy into shaft power to drive the compressor and the engine accessories. A
reduction in pressure ratio of the expanding gases also takes place in the turbine.

In essence, a turbine has just the opposite function to a compressor which performs
work on the fluid to increase the pressure on it. The basic purpose of the turbine is to
extract as much work as possible from the fluid so as to provide acceptable levels of
power drive the compressor. There are many similarities between the design and
features of turbines and compressors. Turbines, like compressors, can be classified as
axial, radial or mixed according to the type of fluid flow being used. Axial flow
turbines work on similar principles as axial flow compressors.

12
According to Saravanamuttoo et al. (2001, pp. 366-7), in the radial flow turbine, fluid
flow with high tangential velocity is directed inwards and leaves the rotor with as
small a whirl (almost axial) velocity as practicable near the axis of rotation. Though
the radial flow turbine has similar components as the centrifugal compressor, a ring of
nozzle vanes replace the diffuser vanes. According to Gorla and Khan (2003, pp. 307-
8), a diffuser is usually used at the exit of the turbine to reduce velocity of exhaust
gases to a negligible value.

According to Håll (2002b, p. 6-8), since the turbine is at the hot end of the gas turbine
engine and is expected to run at high speeds, its blades undergo a number of stresses
such as centrifugal stress, bending stresses, creep, mechanical and thermal fatigue
which can lead to mechanical failure of the blades. In order to ensure durability of the
turbine blades in microturbines, the temperature of the expanding gases (Turbine Inlet
Temperature T3) is limited to a ‘metallurgical limit’. Pullen et al. (2002, p. 91) state
that both radial and axial tip clearances need to be controlled in radial flow turbines
leading to mechanical design failure.

Energy Nexus Group (2002, p. 2) states that in order to obtain optimum overall
efficiency, it is advantageous to operate the expansion turbine at the highest practical
temperature consistent with economic materials and to operate the compressor with
inlet airflow at as low a temperature as possible. According to Dr. J.L. Hill (personal
communication, November 9, 2005), due to advancement in materials technology, it
is possible to increase the turbine inlet temperature, compression pressure ratio and
the efficiency of the microturbine by using high-temperature alloys such as Inconels,
Monels and Nimonics. The main barrier to large-scale use of these materials in
microturbines is the high cost involved.

Gorla and Khan (2003, pp. 307-8) and Pullen et al. (2002, p. 91) assert that the main
advantages of radial flow turbines are that they are compact and can run at high
speeds (to extract greater amount of shaft power). They are particularly efficient when
dealing with small mass flows (as in microturbines) though for larger mass flows,
axial flow turbines are preferable. A single stage in a radial flow turbine can accept a
high-pressure ratio. Most of the reasons for radial flow turbines in microturbines are
the same as those for centrifugal compressors.

13
According to Dambach et al. (2002, pp. 101, 106), another reason for using single-
stage radial flow turbines in microturbines is that they suffer less from tip-leakage
loss as compared to equivalent (usually multi-stage) axial flow turbines. The tip-
leakage loss for an axial flow turbine can be about four times greater than that for a
centrifugal turbine. The tip speed for a centrifugal turbine is significantly greater than
that for an equivalent axial flow turbine. Higher tip speeds enable larger work in
single-stage radial flow turbines than is obtained in single-stage axial flow turbines
and reduce tip leakage and clearance losses.

Pullen et al. (2002, p. 91) suggest the use of mixed flow turbines in microturbine
generators as they are capable of higher mass flow rates than their radial counterparts
for the same expansion pressure ratio (p3/p4) and due to their increased compactness
and efficiency.

Saravanamuttoo (2001, p. 7) states that when operational flexibility is required in the


gas turbine, an additional free (or power) turbine may be used in order to enhance the
efficiency of the system. In the opinion of the author, although a power turbine may
be suitable for larger gas turbines, it is undesirable to use it in a microturbine engine
as it would increase the size and cost of the unit significantly without any appreciable
change in efficiency.

2.3 COMBUSTION CHAMBER

The fuel (either in its original form or after gasification in the gasifier) undergoes
combustion (in the presence of air supplied by the compressor) in the combustion
chamber. Combustion in gas turbines is a continuous process initiated by an electric
spark, and thereafter continued by the self-sustaining flame. A number of types and
configurations of combustion chambers are used in gas turbine engines depending
upon the requirements of the engine (Saravanamuttoo et al., 2001, p. 266).

14
Due to the high temperatures involved in the combustion of fuels, the combustion
chamber has to be highly resistant to mechanical and thermal breakdown. In recent
times, combustion efficiency of the chamber has gained great importance as efficient
combustion of the fuel reduces pollutants such as oxides of carbon (CO2 and CO),
nitrogen (NOx) and unburned hydrocarbons (UHC). Hence, the current goal in
combustion chamber design is to utilise as much of the given fuel as possible
(complete combustion), i.e. attainment of almost 100% combustion efficiency
(Saravanamuttoo et al., 2001, pp. 267-9).

With the advent of various bio-fuels and waste gases as fuels and due to the
introduction of stringent environment protection legislations globally, the need for
better combustion chamber design has become imperative. It may be inferred from
Saravanamuttoo et al. (2001, pp. 266-8) and Shepherd (1960, p. 234-6) that annular
and cannular combustion chambers are well-suited for microturbine engines because
of their compactness and simplicity of design. Bullin (2002, p. 5) suggests the use of
catalysts in the combustion process as an alternative to conventional combustion may
boost the overall efficiency of the microturbine.

2.4 HEAT EXCHANGER

According to Shepherd (1960, p. 5), the basic purpose of a heat exchanger is to reduce
fuel supply for a given required temperature rise by utilizing the heat in the exhaust
gases to raise the temperature of the working fluid immediately after compression.
Eastop and Croft (1990, p. 44-45) classify heat exchangers into three broad categories:
recuperative, regenerative and evaporative. Recuperative heat exchangers (or
recuperators) and regenerative heat exchangers (or regenerators) are commonly used
types of equipment for heat transfer in gas turbines.

Cengel and Turner (2005, p. 374) assert that heat exchangers are suitable for gas
turbines only when the turbine exit temperature T4 is greater than the compressor exit
temperature T2 else heat will flow in the reverse direction (to exhaust gases). Since
gas turbines operating at high pressure ratios (usually in the case of large gas turbines)
encounter this situation frequently, heat exchangers are rarely used there. Hence, heat

15
exchangers are more suitable for microturbines which inherently work at low pressure
ratios.

2.4.1 REGENERATOR

According to Eastop and Croft (2002, p. 181-2), in a regenerator, the hot and cold
fluids pass alternately across a matrix of material of considerable specific heat
capacity; the material is heated by the hot fluid and then cooled by the cold fluid,
thereby making the heat transfer process cyclic in nature. The matrix may be kept
stationary or made to rotate through the hot and cold fluids. According to Shepherd
(1960, p. 250-67), though a regenerator is advantageous with respect to power-to-
weight ratio, excessive leakage losses (of working fluid) make them unsuitable for
long-term use in industrial applications like microturbines. Regenerators are difficult
to manufacture due to the machinability issues associated with (especially in the case
of small regenerators, which could be used for microturbines) and brittleness (which
leads to cracking) of the ceramic materials which are usually used.

2.4.2 RECUPERATOR

Shepherd (1960, p. 250-1) describes a recuperator as a heat exchanger in which heat is


transferred across a solid (usually metallic) wall between the hot and cold fluids. In
the case of recuperated microturbines, the function of a recuperator is to extract heat
from the gas turbine exhaust gases in order to preheat the air used in the combustion
process, and thereby reduce the amount of fuel used to reach operating temperature
(Bullin, 2002, p. 3). Stares and Mabbutt (2002, p. 69) state that predominantly
recuperators used in microturbines are “metallic of a matrix or honeycomb-type
construction”. Recuperated microturbines are more popular than their unrecuperated
counterparts in industrial applications.

According to Bullin (2002, p. 8-10), stainless steel primary surface recuperators are a
popular choice for heat exchange in microturbines currently while research is being
carried out on alternative recuperator design utilizing brazed fin and plate technology.
In primary surface recuperators, the plates are not bonded together. They are instead
welded together around the edge, and are clamped together thus allowing for thermal

16
expansion without the transmission of high stresses to the joints (as is the case with
brazed recuperators). This tends to increase reliability and durability due to reduced
probability of thermal stress failure and consequential leakage of fluids. It can further
be inferred that the box-shaped recuperator is a well-suited type of recuperator for
microturbines due to its high effectiveness and comparatively low capital costs.

According to Stares and Mabbutt (2002, p. 69), chief design constraints for
recuperators are the need for cost-effectiveness and high thermal effectiveness ε.
Long-term recuperator performance is dependent on the type of fuel and thermal
cycling history. Since recuperators work at the hot end of the microturbine, efforts are
being made to reduce the formation of thermal fatigue (leading to mechanical failure)
in the recuperator material. Since recuperators are usually comprise of robust
materials such as stainless steels and high temperature alloys, they are usually
preferred to regenerators (which use brittle materials such as ceramic) for use in
microturbines.

According to Dr. J.L. Hill (personal communication, November 9, 2005), a major


problem faced by recuperators in microturbines running on refuse-based fuels or
biomass is the formation of deposits of unburned particulate matter on the honeycomb
structure, thereby leading to accelerated corrosion of the recuperator material and
reduced effectiveness of the recuperator. High temperatures at the hot end further
contribute to this corrosion process. Hence, efficient combustion of the fuels is an
essentiality for efficient recuperation of heat from exhaust gases. High temperature
alloys are well-suited materials for recuperators but their cost-effectiveness currently
makes them impractical for large-scale manufacture of microturbine generators.
Borbely and Kreider (2001, p. 280) state that the use of recuperators in microturbines
limits the amount of recovered heat which could be used for CHP applications such as
water heating and space heating.

17
2.5 OTHER COMPONENTS

According to Energy Nexus Group (2002, p. 5), single-shaft microturbines contain


digital power controllers to convert the high frequency Alternating Current produced
by the electric generator into usable electricity. The high frequency Alternating
Current is rectified to Direct Current, inverted back to 60 or 50 Hz AC, and then
filtered to reduce harmonic distortion. This is a critical component in the single-shaft
microturbine design and represents significant design challenges, specifically in
matching turbine output to the required load. Most microturbine power electronics are
generating 3-phase electricity.

The shaft of the microturbine is typically supported on two bearings. According to


Pullen et al. (2002, p. 93-4), although conventional oil bearings are extremely durable,
they are only suitable if the shaft has a low diameter and large clearances to reduce
excessive losses. Leakage of oil is another problem associated with this type of
bearing. Magnetic bearings and air bearings are new alternatives but are cost-effective.

According to Saravanamuttoo (2001, p. 36-7), a gasifier can be attached to a


microturbine when the fuel (solid or liquid) being used contains high levels of
impurities (usually vanadium and sodium, which may cause corrosion of the turbine).
The purpose of the gasifier is to transform a fuel such as poor-quality coal or heavy
oil into a clean gaseous fuel. The usage of gasifier along with a microturbine plant
gives potential for power generation using virtually any waste material (such as solid
refuse or liquid slurry).

The nozzle design (convergent, divergent or convergent-divergent) and performance


also affects the overall performance of the microturbine system to some extent as
choking losses occurring in the nozzle reduce thermal efficiency of the system.

18
3. CYCLE ANALYSIS

The basic working principles of a microturbine are the same as those for gas turbines.
Both are forms of heat engines which can be simply described using the First and the
Second Law of Thermodynamics. Gas turbines are basically engines which convert
chemical energy into mechanical energy and finally into electrical and heat energy (in
Combined Heat and Power installations). The following sections contain a
thermodynamic analysis of the Brayton Cycle, both unrecuperated and recuperated.
The key difference between the ideal analyses and the actual analyses is the fact that
the former does not take into account efficiencies of the components. Other
modifications such as reheating and intercooling can be made to the cycle but these
modifications will not be discussed in the study as they are not used by microturbines
currently.

3.1 IDEAL SIMPLE BRAYTON CYCLE

According to the Cengel and Turner (2005, pp. 367-8), the ideal Brayton Cycle is a
gas power cycle which is composed of four internally reversible processes:

Process 1-2 Isentropic compression


Process 2-3 Isobaric heat addition
Process 3-4 Isentropic expansion
Process 4-1 Isobaric heat rejection

In the analysis of the ideal simple cycle, the following air-standard assumptions are
being made (Saravanamuttoo et al., 2001, pp. 45-6; Shepherd, 1960, p. 27):

a) Processes 1-2 (compression) and 3-4 (expansion) are both adiabatic (processes
involving no heat transfer) and reversible in nature.

b) There are no losses in any of the components during the thermodynamic cycle.

19
c) The properties of the working fluid (such as density and molecular weight) used in
the cycle remain constant throughout the thermodynamic cycle, thereby
suggesting that the working fluid is a perfect gas.

d) The specific heat capacities at constant pressure (cP) and at constant volume (cV)
and the heat capacity ratio (γ) of the working fluid are constant throughout the
thermodynamic cycle.

e) Complete (perfect) combustion of the fuel is taking place in the combustion


chamber.

f) No leakage occurs in the system, thereby emphasising that the mass flow of the
working fluid is constant throughout the cycle.

g) The mass of the fuel is negligible.

h) The compressor inlet pressure is equal to the turbine exit pressure, thereby
emphasizing that there are no pressure losses during the course of the cycle.

Using the Steady Flow Energy Equation for a single-stream system, the cycle
efficiency of the Brayton Cycle can be obtained. Due to the assumptions being taken,
the energy balance for the system (Cengel and Turner, 2005, pp. 542-3) may be
written as follows:



1 2
 
Q in  Wout  m .he  hi   . ue  ui  g .ze  zi 
2
… (1)
 2 

Q in is the rate of heat supply to the system

Wout is the rate of work done (shaft work, in this case) by the working fluid

m is the mass flow rate of the working fluid

20
hi and he are the enthalpies b of the working fluid at the inlet and exit states
respectively
ui and ue are the average flow velocities of the working fluid at the inlet and exit states
respectively
g is the acceleration due to gravity
zi and ze are the elevations of the inlet and exit states

If equation (1) is divided by the mass flow rate m , the following expression will be
obtained:

Q in W out
m

m
1
2
 2

 he  hi   . u e  ui  g .z e  zi 
2
… (2)

In the case of an ideal gas turbine, it may be assumed that the difference between the
elevations of the inlet and exit points of the working fluid is negligible, i.e. ∆z ≈ 0,
and that the change in kinetic energy of the working fluid between the inlet and is
negligible, which implies that ∆u ≈ 0 (since it has already been stated that the mass
flow remains constant throughout the cycle in assumption ‘f’ above). If Q in m  Q

and Wout m  W , equation (2) may be modified as follows:

Qin  Wout  he  hi  … (3)


(derived from Cengel and Turner, 2005, pp. 542-3)

It can be clearly seen that equation (3) is a form of the ‘First Law of
Thermodynamics’ according to which the “total energy of the system during a process
is equal to the difference between the total energy entering the system and the total
energy leaving the system during the process” (Cengel and Turner, 2005, p. 159).

b
All work (W) and heat (Q) quantities used in the calculations have been expressed as energy divided
by mass flow of air. Another assumption which is being made throughout the analysis is that the mass
 f ) is negligible as compared to the mass flow rate of air ( m a ) used in the
flow rate of the fuel ( m
system. Hence, it may be assumed that m a  m f   m a .

21
If equation (3) is applied to the various processes of the ideal Brayton Cycle, the
following expressions will be obtained (Cengel and Turner, 2005, p. 368;
Saravanamuttoo et al., p. 46):

W12  h1  h2  WC Work done by compressor on working fluid … (4)

Q23  h3  h2  Qin Heat supplied to system … (5)

W34  h3  h4  WT Work done on turbine by working fluid … (6)

Q41  h4  h1  Qout Heat emitted from system … (7)

For any thermodynamic cycle, the thermal efficiency ηth can be described as:

net work output Wnet


th   … (8)
heat supplied Qin
(Cengel and Turner, 2005, p. 227)

In the case of a Brayton Cycle, the net work done by the system (Wnet) is the sum of
the work done on the compressor (WC) and the work done by the turbine (WT):

Wnet  W34  W12  h3  h4   h1  h2  … (9)


(Saravanamuttoo et al., 2001, p. 46)

By substitution of Wnet and Qin from equations (9) and (5) into equation (8), the
thermal efficiency of the Brayton Cycle may be obtained as:

th , Brayton 
h3  h4   h1  h2 
h3  h2
(Saravanamuttoo et al., 2001, p. 46)

22
Since h  c P .T for a gas in equilibriumc, the above equation (Saravanamuttoo et
al., 2001, p. 47) may be re-arranged as follows:

c P .T3  T4   c P .T1  T2 
th , Brayton 
c P .T3  T2 

Since it is being assumed that the specific heat capacity at constant pressure of the
working fluid remains constant, the above equation may be arranged as follows:

T3  T4  T2  T1
th , Brayton  … (10)
T3  T2

Since processes 1-2 (compression) and 3-4 (expansion) are assumed to be isentropic,
it can be said that (Saravanamuttoo et al., 2001, p. 47):

  1    1 
    
T2  p2     T  p    
   and 3   3 
T1  p1  T4  p 4 

Since pressure losses are negligible, pressures at various points of the cycle may be
related as p2 p1  p3 p 4  constant. Here, the constant may be represented by Π and
is known as the pressure ratio.

p2 p
 3  … (11)
p1 p4
  1 
 
T2 T3  
   … (12)
T1 T4
(Saravanamuttoo et al., 2001, p. 47; Cengel and Turner, 2005, p. 368)

  1    1 
     
Therefore, T2  T1 .  
and T3  T4 .  

c
All temperatures used for calculations are in K (absolute) although in actual practise, °C is a popular
unit for temperature. K = °C + 273

23
Substitute T2 and T4 in equation (10):

1
th ,Brayton  1    1 
… (13)
 
  

(Hodge, 1955, p. 160; Cengel and Turner, 2005, p. 368)

In equation (13), it may be noted that the ideal Brayton Cycle efficiency is
independent of inlet or exit temperatures. It is only dependent on two factors, namely
the pressure ratio of the system (Π) and the heat capacity ratio (γ) of the working fluid
(which represents the atomicity of the working fluid as mono-atomic, di-atomic, etc.).
Therefore, in an ideal scenario, if the pressure ratio of the system is increased, the
thermal efficiency of the system increases. An interesting point which may be noted is
that if the working fluid is compressed to an infinite pressure ratio, the system tends to
become 100% efficient, i.e. lim th ,Brayton   1 .
 

The values for the Ideal Brayton Cycle Efficiency for pressure ratios of 1 to 40 have
been calculated and plotted in Graph (1). The value of γ is taken to be 1.4 for dry air.
It can be seen that the value of efficiency keeps increasing continuously along with
the pressure ratio. It may also be noted that for a unity pressure ratio (which implies
that no compression of the working fluid is taking place); the efficiency of the cycle is
zero, which means that no positive work output can be obtained from the gas turbine.

Another significant quantity which affects the feasibility of a gas turbine is the
Specific Work Output. It is a major factor deciding the size of a gas turbine power
plant for a given power. The specific work output can be determined by modifying
equation (9) for net work output into a non-dimensional form as follows
(Saravanamuttoo et al., 2001, p. 47):

Wnet  T3 T4   T2 
     1  
c P .T1  T1 T1   T1 

24
  1 
 
   T3
T2  T1 . and T4    1 
may be substituted into the expression above:
 
  

   
  1 
 
Wnet  T3 1 T3  T1 .    
   .   1    1  
c P .T1  T1 T1  
  T1 
     

  1   
Wnet  T3 
  
  
1 
  . 1    1   … (14)
cP .T1  T1  
      
 

(derived from Saravanamuttoo et al., 2001, p. 47; Hodge, 1955, p. 160)

In equation (14), it may be observed that the Specific Work Output of the cycle is
dependent on the Compressor Inlet Temperature (T1) and Turbine Inlet Temperature
(T3) along with pressure ratio and heat capacity ratio unlike the cycle efficiency that is
independent of temperature. Since T1 is the ambient temperature and is not under
human control, the only factors which can susceptible to human intervention in this
case are the pressure ratio (Π) and the turbine inlet temperature (T3).

In practice though, T3 can only be increased up to a certain limit as the metallic


turbine blades, being at the hot end of the gas turbine, undergo mechanical failure due
to the generation of thermal and mechanical ‘creep’ (both of which complement each
other) by the high temperatures present in the turbine region of the gas turbine. This
limit on the turbine inlet temperature is known as the ‘metallurgical limit’ and has
affected the demand of gas turbines and microturbines as power generation equipment
adversely. Research is being carried out in order to manufacture turbine blades with
high-temperature alloys which can tolerate high turbine inlet temperatures as higher
values of T3 will increase the specific work output from the gas turbine substantially
(Saravanamuttoo et al., 2001, pp. 47-8).

25
Currently, excessive costs of high-temperature alloys act as a barrier to the
manufacture of turbine blades for microturbines from these materials though there is a
future possibility of such a revolution taking place once the benefits related to use of
high-temperature alloys for microturbine blades exceed their costs and the demand for
microturbines increases.

Graph (2) shows the values of Specific Work Output under the following conditions:

γ = 1.4 (for dry air)


T1  15°C (288K)
T3 from 500°C (773K) to 1500°C (1773K) at increments of 50°C
 from 1 to 40 at increments of 1

The pressure ratio (Πopt) for optimum specific work output for a given compressor
inlet temperature and turbine inlet temperature can be calculated by differentiating
equation (16) with respect to Π as follows:

     1  
  Wnet    T3  1     
   
   . 1    1  
 1   0
  cP .T1  opt   T1   
  
       

 
T3  1      
  1 

.   opt 0
T1    1    
     
 opt 

2
   1   T
  
 opt   3
  T1

  
 
T   2  1  
 opt   3  … (15)
 T1 
(Saravanamuttoo et al., 2001, p. 47-8)

26
In order to obtain the optimum specific value for any given values of T1 and T3, Πopt
can be substituted from equation (15) into equation (14):

 
 
1    T3  
1/ 2
 Wnet  T3 
   .1  1/ 2 
 1    
 c P .T1  opt T1
  3     1  
 T  T
  T1  

2
 Wnet   T 
    3  1 … (16)
 c P .T1  opt  T1 
(derived from Saravanamuttoo et al., 2001, p. 48)

The optimum values of pressure ratios and corresponding specific work outputs have
been tabulated in represented by a straight line on Graph (2). The value of γ = 1.4 for
dry air has been used. The reason for the decrease in specific work output after an
increase is that at higher compression pressures, the effect of the pressure ratio
dominates over the effect of the temperature ratio (ratio of T3 and T1). According to
Cengel and Turner (2005, p. 39), this leads to a need for compromise between specific
work output and efficiency. If the specific work output is low then the larger mass
flow (leading to a larger gas turbine unit) is required to obtain the same power output.

27
3.2 IDEAL BRAYTON CYCLE WITH HEAT EXCHANGE

This is a modification of the simple Brayton Cycle where an ideal heat exchanger
(with an effectiveness of 100%) is added to the system. In this case, T5 and T6 are
equal to equal to T4 and T2 respectively. Therefore, exhaust heat from process 4-6 can
be completely transferred to process 2-5 after compression (1-2), thereby conserving
the fuel which would have otherwise been required to raise the temperature of the
working fluid from T2 to T4. The net work output is not affected by the addition of a
heat exchanger but the heat supplied to the system (Qin) reduces from h3  h2 to

h3  h5 and the heat emitted by the system (Qout) reduces from h4  h1 to

h6  h1 (Saravanamuttoo et al., 2001, pp. 48-9; Cengel and Turner, 2005, 374-5).

In addition to the assumptions stated for the simple Brayton Cycle earlier, an
additional assumption applicable to the heat exchange cycle is that the heat exchanger
used in the system in a countercurrent flow heat exchanger which performs ‘complete
heat exchange’. So, the “temperature rise on the cold side is maximum possible and
exactly equal to the temperature drop on the hot side” (Saravanamuttoo et al., 2001, p.
46; Cengel and Turner, 2005, p. 374-5).

Using equations (5), (8) and (9), the thermal efficiency of the ideal Brayton Cycle
with Heat Exchange will be:

th , HE 
h3  h4   h1  h2  … (17)
h3  h5 
(Saravanamuttoo et al., 2001, p. 48)

Here, h5 can be obtained from the expression for the effectiveness (ε) of a heat
exchanger which is:

h5  h2

h4  h2
(Cengel and Turner, 2005, p. 375)

28
h5   .h4  h2   h2 … (18)

Substitute h5 into from equation (18) into equation (17):

h3  h4   h1  h2 
 th , HE  … (19)
h3   .h 4  h 2   h2 

Since an ideal heat exchanger is being used in this case, ε = 100%:

th , HE 
h3  h4   h1  h2 
h3  h4
(derived from Saravanamuttoo, 2001, p. 48)

Since h  c P .T for a gas in equilibrium, the above equation may be re-arranged as
follows:

c P .T3  T4   c P .T1  T2 
th , HE  , or
c P .T3  T4 

T2  T1
th , HE  1 
T3  T4

  1 
   T3
Since the processes are isentropic, T2  T1 .  
and T4    1 
can be
 
  

substituted into the above equation:

  1 
 
  
T1 .  T1
th , HE  1 
T3
T3    1 
 
  

29
  1 
  
  
th , HE  1  … (20)
T3
T1
(Saravanamuttoo et al., 2001, p. 48-9; Cengel and Turner, 2005, p. 375)

From equation (20) and Graph (3), it may be deduced that th, HE reduces if there is an

increase in pressure ratio and increases if there is an increase in the turbine inlet
temperature. Contrary to the efficiency of the simple cycle, the efficiency of the heat
exchange cycle decreases with an increase in pressure ratio after an initial increase.
Similar to specific work output, thermal efficiency of the heat exchange cycle also
increases up to an optimum value and then decreases due to the dominance of the
pressure ratio over the temperature ratio. According to Saravanamuttoo (2001, p. 49),
1  
at higher pressures than   T3 / T1 2   1  [equation (15)], the heat exchanger starts
. 

to cool the air leaving the compressor, thus reducing the overall thermal efficiency.

Hence, it may be inferred that a heat exchange Brayton Cycle is most efficient at low
pressure ratios and reduces the fuel consumption of the gas turbine, thereby making it
a good substitute to the simple Brayton Cycle at low pressure ratios.

The values of efficiency of the ideal Brayton Cycle with Heat Exchange for the
following parameters have been plotted on Graph (3):

ε = 100%
γ = 1.4 (for dry air)
T1  15°C (288K)
T3 from 500°C (773K) to 1500°C (1773K) at increments of 50°C
 from 1 to 40 at increments of 1

As the net work output for a heat exchange cycle is the same as the simple cycle, the
specific work output also remains the same [equation (16)]. Hence, Graph (2) for
specific work output applies to the heat exchange cycle as well.

30
3.3 ACTUAL SIMPLE BRAYTON CYCLE

The actual Brayton Cycle has much lower efficiencies than the ideal cycle because it
uses real components (compressor and turbine) which have efficiencies lower than
100% due to the presence of number of component losses such as friction and
turbulence. Most of the assumptions stated for ideal systems earlier do not apply to
actual systems. Other combustion losses (due to incomplete combustion of fuel),
mechanical losses (due to bearings, auxiliaries, etc.) and loss of heat to surroundings
further reduce the efficiency of the cycle. Leakage losses can reduce the mass flow
rate of air and fuel considerably. The main reason for low actual cycle efficiencies is
that under actual conditions, the thermodynamic processes become irreversible.
Pressure losses in the system during constant pressure (ideally) heat addition and heat
rejection processes also contribute to these discrepancies. Other types of losses may
also occur in the system but these are usually ignored (Saravanamuttoo et al., 2001,
pp. 53-66; Shepherd, p. 27, 47-49). The changes in ideal assumptions have been listed
below:

a) All processes are irreversible in nature as some amount of energy is lost to the
environment.

b) Components losses are being considered during the thermodynamic cycle. Hence,
the efficiencies of the compressor (ηC), turbine (ηT) and other components are not
100%. Also, mechanical efficiency (ηm) is less than 100%.

c) The properties of the working fluid used in the cycle do change throughout the
thermodynamic cycle, but they are being assumed constant during the actual cycle
calculations.

d) Different values for constants such as specific heat capacity at constant pressure
(cP) and the heat capacity ratio (γ) are being considered for air during compression
processes 1-2a and for combustion gases during expansion processes 3-4a. For
more accurate results, these values are expected to vary with temperature.
Dry air: cP,a = 1.005 kJ/kg.K, γa = 1.40
Combustion gases: cP,g = 1.148 kJ/kg.K, γg = 1.33

31
e) Incomplete combustion of the fuel is takes place in the combustion chamber with
an efficiency ηB.

For the actual cycles, the letter ‘a’ will be used to differentiate between actual and
isentropic parameters. The actual Compressor Exit Temperature (T2a) is higher than
the isentropic Compressor Exit Temperature (T2) and the actual Turbine Exit
Temperature (T4a) is higher than the isentropic Turbine Exit Temperature (T4). Due to
these changes in the turbine exit temperature and compressor exit temperature, the
area under the curve reduces from Area (1-2-3-4-1) to Area (1-2a-3-4a-1). Since the
area under the T-s curve represents net work done by the system, this reduction in
area (under the curve) reduces the net output and thermal efficiency of the gas turbine
system (Thomson, 1949, pp. 39-41).

Now, the energy equations (4), (5), (6) and (7) will change to:

W12 a  h1  h2 a Work done by compressor on working fluid … (21)

Q2 a 3  h3  h2 a Heat supplied to system … (22)

W34 a  h3  h4 a  WT ,a Work done on turbine by working fluid … (23)

Q4 a 1  h4 a  h1  Qout ,a Heat emitted from system … (24)

Mechanical losses occur during transmission of power mainly due to bearing friction
and windage in the ducts. So, the net work done by the compressor changes to
WC ,a  h1  h2 a  m where WC ,a is actual turbine work which is required to drive the

compressor divided by mass flow and m is the mechanical transmission efficiency


for transmission of power from the turbine to the compressor (Saravanamuttoo et al.,
2001, p. 66). An assumed value of  m  99% has been used in the calculations.

32
Now, the net work done by the system (Wnet) [equation (9)] will change to:

Wnet ,a  WT ,a  WC ,a  h3  h4 a  
h1  h2 a  … (25)
m
(derived from Saravanamuttoo et al., 2001, p. 75)

Similarly, due to incomplete combustion of the fuel in the combustion chamber, the
heat needed by the system increases and the combustion efficiency (ηB) is lower than
100%. Therefore, the heat supplied by the fuel becomes:

h3  h2 a
Qin ,a  … (26)
B

An assumed value of  B  98% has been used in the calculations.

Fixed values for Specific Heat Capacity (cP) and Heat Capacity Ratio (γ) have been
used in the design calculations but in reality, they vary with temperature changes and
changes in chemical composition of the working fluid due to internal combustion,
though these effects are usually negligible (Saravanamuttoo et al., 2001, p. 66-8). For
more accurate results, property tables or charts for gases may be used to obtain
enthalpy values at each temperature.

Equation (11) for pressure ratio is valid only for the ideal Brayton cycle. According to

Saravanamuttoo et al. (2001, p. 61-3), in the actual cycle, p2  p3 due to the


p1 p4
presence of pressure losses in the various components and ducting. In the combustion
chamber, a loss in stagnation pressure pb  occurs due to the aerodynamic resistance

of flame-stabilizing and mixing devices. An assumed value of pb  2% compressor


delivery pressure has been used in the calculations. The turbine inlet pressure changes
 p 
from p3  p2 (ideal case) to p3a  p2 .1  b  .
 p2 

33
The turbine exit pressure remains the same, i.e. p4 a  p4 . So the pressure ratio at
expansion changes to:

p3a p2  pb 
'   . 1
p4 a p4  p2 

p2
Since p4 = p1 and   , the equation may be re-modelled as follows:
p1

 pb 
 '  .1   … (27)
 . p1 
(derived from Saravanamuttoo et al., 2001, p. 61-3)

The expressionsd for Compressor Efficiency ( C ) and Turbine Efficiency ( T ) are as


follows:

WC h  h2 c .(T  T2 )
C   1  P ,12 1 … (28)
WC ,a h1  h2 a c P ,12 a .T1  T2 a 

WT ,a h3  h4 a c P ,34 a .(T3  T4 a )
T    … (29)
WT h3  h4 c P ,34 .T3  T4 

(Cengel and Turner, 2005, p. 373; Saravanamuttoo et al., 2001, p. 56)

For actual cycles, the specific heat capacity changes according to the temperature. In
equations (28) and (29), the numbers in the subscript of cP represent the process for
which the energy is being calculated.

d
Isentropic efficiencies are being used in this case although polytropic efficiencies may be used when
dealing with multi-stage turbomachinery.

34
By assuming c P ,12  c P ,1 2 a  c P ,a and    a (i.e. constant c P and  are being

assumed for the compression processes) in equation (29) and substituting


  1 
  
T2  T1 .  
from equation (12), actual Compressor Exit Temperature can be

determined. c P ,a and  a are the specific heat capacity and heat capacity ratio for dry

air during the compression process:

   a 1 
  
   a   1 
T2 a  T1 . 1   … (30)

C 
 
(Saravanamuttoo et al., 2001, p. 57)

By assuming c P ,34  c P ,3 4 a  c P ,g and    g (i.e. constant c P and  are being

assumed for the expansion processes) in equation (29) and substituting


  1 
T4  T3 /  ' 
 
  from equation (12) and (27), actual Turbine Exit Temperature can

be determined. c P , g and  g are the specific heat capacity and heat capacity ratio for

exhaust gases during the expansion process.

 
 
 
T
T4 a  T3 . 1   T   … (31)
   g 1 
  
    p b    g  
  . 1  

    . p1   

(Saravanamuttoo et al., 2001, p. 57)

Now, equations (25) and (26) can be re-modelled as follows:

c P , g .T3  c P ,a .T2 a
Qin ,a  … (32)
B
(Cengel and Turner, 2005, p. 368)

35
c P ,a T1  T2 a 
Wnet ,a  c P ,g T3  T4 a   … (33)
m
(Saravanamuttoo et al., 2001, p. 75)

Substitute T2a and T4a from equations (29) and (30) in equations (32) and (33)
respectively:

   a 1 
  
1     a   1  
Qin ,a  .c P , g .T3  c P ,a .T1 .1   … (34)
B    
 C
  
(derived from Cengel and Turner, 2005, p. 368)

   1g  
      a 1  
  
    pb   g   1
   1    a  
Wnet ,a  c P ,g .T3 .T .1  .1     c P ,a .T1 .  … (35)
   . p1    m   C 

     
(derived from Saravanamuttoo et al., 2001, p. 47; Cengel and Turner, 2005, p. 368)

Substitute Wnet ,a and Qin ,a in equation (8) in order to determine the actual cycle

efficiency:

   
 g 1     
 a 1 

T3    pb     g   
c P , a 1    a  
c P , g . .T .1  .1     . 
T1    . p1     m  C 
   
th ,Brayton ,a  … (36)
  

 a 1
 

1   T3   1    a  
.c .   c .1  
 B  P , g  T1  P ,a  C 

  
(derived from Saravanamuttoo et al., 2001, p. 47; Cengel and Turner, 2005, p. 368)

According to Hodge (1955, pp. 102-3), if the compressor efficiency and the turbine
efficiencies for a simple cycle are numerically equal then the turbine efficiency would
have a greater effect on the thermal efficiency and specific work output of the cycle.
This occurrence can be accounted for by the fact that “the whole expansion work (WT)

36
is affected by turbine efficiency (ηT) while the compressor efficiency (ηC) only affects
the results of the work (WC) used internally, which, for all positive values of overall
efficiency, is a smaller quantity”.

Graphs (4a-f) describe the effect of changing compressor efficiencies on the thermal
efficiency of the actual Brayton Cycle. The following parameters have been used in
the graphs:

γa = 1.4 (for dry air)


γg = 1.33 (for combustion gases)
cP,a = 1.005 kJ/kg.K (for dry air)
cP,g = 1.148 kJ/kg.K (for combustion gases)
T1  15°C (288K)
T3 from 500°C (773K) to 1500°C (1773K) at increments of 50°C
 from 1 to 40 at increments of 1
ηC from 70% to 95% at increments of 5%
ηT = 85%
ηB = 98%
ηm = 99%
p1 = 1 bar
∆pb = 2% compressor delivery pressure

In each of these graphs, it can be observed that thermal efficiency increases with
increasing turbine inlet temperatures. If each of the graphs is observed, it can be
inferred that increasing compressor efficiencies increase the thermal efficiency of the
cycle. If cycle efficiencies are compared in Graphs (4a) and (4f) at turbine inlet
temperature 1500°C and at a pressure ratio of 4:1, it can be seen that an increase of
25% in the compressor efficiency increases the overall efficiency by only 2.66%. In
Graph (4f), it can be observed that for the given parameters, the thermal efficiency at
turbine inlet temperature of 1500°C (and at a pressure ratio of 4:1) is 6.57% higher
than thermal efficiency at 500°C. Since these differences are not very large, it may be
inferred that at low pressure ratios, the thermal efficiency of the cycle is not affected

37
significantly by increasing turbine inlet temperatures and thus, unrecuperated are not
viable for operation at low pressure ratios.

Graphs (5a-f) show the effect of changing turbine efficiencies on the thermal
efficiency of the cycle of the Brayton Cycle. The same parameters have been used as
for Graphs (4a-f) except the compressor and turbine efficiencies. In this case:

ηC = 85%
ηT from 70% to 95% at increments of 5%

The observations with respect to the thermal efficiency and turbine inlet temperatures
are similar to those for Graphs (4a-f). While comparing Graphs (4f) and (5f), it may
be observed that the thermal efficiencies at T3 = 1500°C and Π = 4, the efficiency in
the former case is 21.53% and in the latter case is 24.32%. This shows that the effect
of the turbine efficiency on the cycle efficiency is greater than the effect of the
compressor efficiency. Hence, methods to improve turbine efficiency need to be
researched to achieve greater cycle efficiencies for the microturbine.

If cP = cP,g is assumed for the complete cycle then according to equations (14) and
(35), the specific work output will become:

    
 g 1     
 a 1 

Wnet ,a T .    p b     g   c P , a  1    a  
 3 T .1  .1    .  … (37)
c P ,g .T1 T1    . p1    c P ,g   m .C 
   
(derived from Saravanamuttoo et al., 2001, p. 47; Hodge, 1955, p. 93)

38
3.4 ACTUAL BRAYTON CYCLE WITH HEAT EXCHANGE

The actual Brayton Cycle with Heat Exchange is basically a modification of the actual
Brayton Cycle (simple) which includes a real heat exchanger [with effectiveness (ε)
less than 100%]. In this case, only a part of the exhaust heat is transferred from
process 4a - 6 to process 2a - 5 as some of the energy is dissipated from the system
during heat transfer.

The net work output (Wnet,a) will remain the almost the same as that in the simple
actual cycle with a small difference because of additional pressure losses in the heat
exchanger. Since a real heat exchanger is being included in the system, pressure
losses [other than stagnation pressure losses (Δpb) in the combustion chamber] occur
in it due to friction in the passages on the air-side (Δpha) and the gas-side (Δphg) of the
heat exchanger (Saravanamuttoo et al., 2001, p. 61-3). Assumed values of pha  3%

compressor delivery pressure and phg  0.04 bar have been used in the calculations.

 p p 
The turbine inlet pressure changes from p3  p2 to p3a ,hex  p2 .1  b  ha  .
 p2 p2 

The turbine exit pressure changes from p4  p1 to p4 a ,hex  p1  p hg . So, the

pressure ratio at expansion becomes:

 pb p 
.1   ha 
p3a ,hex . p1 . p1 
' '    … (38)
p4 a ,hex p1  phg

Now, T2a remains the same as that in equation (32) but T4a changes to:

 
T
T4 a ,hex  T3 . 1  T  
  … (39)
  
 g 1 
 ' '  g  


(derived from Saravanamuttoo et al., 2001, p. 75)

39
Here, h5 can be obtained from the expression for the effectiveness (ε) of a real heat
exchanger which is:

h5  h2 a
 … (40)
h4 a ,hex  h2 a

(Cengel and Turner, 2005, p. 375)

h5   .h4 a ,hex  h2 a   h2 a … (41)

Since, heat energy supplied to the system (Qin,a) is h3 – h5, it can be expressed as:

h3   .h4 a ,hex  h2 a   h2 a 
Qin ,a ,hex  … (42)
B

Since h  c P .T , equation (45) can be re-modelled as follows:

c P , g .T3   .T4 a ,hex   c P ,a .T2 a .1   


Qin ,a ,hex  … (43)
B

If equation (43) is subtracted from equation (32), it can be observed that the heat
supplied to a heat exchange cycle (Qin,a,hex) reduces by an amount Qin,a,red as compared
to heat supplied to the simple cycle (Qin,a) such that:


Qin ,a ,red  cP,g .T4a ,hex  cP,a .T2 a  … (44)
B

Substitute T2a and T4a,hex from equations (30) and (39) into equations (33) and (42) to
obtain Wnet,a and Qin,a:

c P ,a .T1   a  


 a 1 
  
 g 1 
 c P , g .T3 .T .1   ' '   g     1
 
Wnet ,a ,hex .  … (45)
   m .C  

(derived from Saravanamuttoo et al., 2001, p. 75)

40
  
 a 1 

cP , g .T3    g 1 
 c P ,a .T1  1    a  
.1      .T .1   ' '  .1   
 
Qin ,a ,hex   g    .1 
B    B  C 

 
… (46)
(derived from Saravanamuttoo et al., 2001, p. 75; Cengel and Turner, p. 368)

Substitute Wnet,a and Qin,a from equations (45) and (46) into equation (8) to obtain the
thermal efficiency for the actual Brayton Cycle with heat exchange:

  
 g 1 


   p b p ha    g  

  .1  . p  . p    c .T   a 1  
c P , g .T3 .T .1    1 1
  P ,a 1 .   a   1

  p1  p hg
   m . C  
    

   
 th ,hex   
   g 1  
   
    p b p    g   
  a 1  

  .1   ha     c .T .1     1    a  
c P , g .T3     . p  . p1  
. 1      .T .1  
 1


 
P ,a 1
.1  
B    p1  p hg
   B  C 
       
     
  
… (47)
(derived from Saravanamuttoo et al., 2001, p. 75; Cengel and Turner, p. 368)

Graphs (6a-f) show the effect of changing heat exchanger effectiveness on the thermal
efficiency of the recuperated Brayton Cycle. Thermal efficiencies have been plotted
using the same parameters as Graphs (4a-f) with a few differences as stated below:

ηC = ηT = 85%
ε from 70% to 95% at increments of 5%

41
In each of these graphs, it can be observed that thermal efficiency increases with
increasing turbine inlet temperature. If each of the graphs is compared, it may be
inferred that increasing heat exchanger effectiveness increases the thermal efficiency
of the cycle. An important observation which can be made from these graphs is that
the peak of each of the curves (at low pressure ratios) rises significantly with
increasing heat exchanger effectiveness but the rest of the curve is virtually unaffected.

If Graphs (6a) and (6f) are compared for Π = 40 and T3 = 1500°C, the cycle efficiency
increases by a mere 0.72% whereas for Π = 4 and the same temperature, it increases
by 19.47%. The reason for this phenomenon is the cooling which takes place in the
system (due to the heat exchanger, as described earlier) at pressure ratios higher than
the optimum pressure ratio. At Π = 4 and T3 = 1500°C, Graphs (5d) and (6d) give
cycle efficiencies of 20.67% and 47.41% respectively. This difference of 26.74% in
the thermal efficiency shows the clear advantage of using a heat exchanger for low
compression pressure ratios of 4:1.

This argument can be further supported by Graph (7), in which at Π = 4 and T3 =


900°C, the simple cycle has a thermal efficiency of 18.02% and that of the
recuperated cycle is 36.37%, i.e., a twofold increase (approximately), thereby
exhibiting the beneficial recuperative effect of heat exchange on the cycle efficiency
at low pressure ratios.

If cP = cP,g is assumed for the complete cycle then according to equations (14) and
(35), the specific work output will become:

  
 g 1 
 
   pb p ha    g  
   a 1  
. 1  
Wnet ,a ,hex T3 .T    . p1 . p1    c    a   1 
 .1      P ,a .  … (48)
c P , g .T1 T1 

p1  phg
  cP,g   m .C 
     
   
 
(derived from Saravanamuttoo et al., 2001, p. 47, 75; Hodge, 1955, p. 113)

42
3.5 ELECTRIC POWER OUTPUT AND SHAFT POWER REQUIRED

The gas turbine produces shaft power. It can be connected to an electric generator in
order to convert this shaft (mechanical) power into electric power. Some amount of
this power is lost during this conversion (mechanical to electrical). Thus, the
efficiency of the electric generator is always lower than 100%.

Suppose the electric power output of a plant is PE, then the shaft power that has to be
produced by the gas turbine has to be higher than the rated electric power (PE) of the
plant due to power losses in the generator. The required shaft power (PM) may be
calculated by using the following expression:

PE
PM  kW … (49)
E

ηE is the efficiency of the electric generator and its value is being assumed to be 95%
for the calculations.

If the number of hours for which the system runs annually is AH, then total energy (ET)
which needs to be generated by the gas turbine in order to work at full load per annum
is:

ET  PM . AH … (50)

43
3.6 FUEL CONSUMPTION AND COSTS

Mass flow rate of air ( m a ) required for a power plant of shaft power PM under the
given conditions can be determined using the following equation:

PM
m a  … (51)
Wnet
(Saravanamuttoo et al., 2001, p. 75)

It may be inferred from Saravanamuttoo et al. (2001, pp. 47-76) that the heat supplied
to the system (simple and heat exchange cycle) is the product of the fuel-air ratio fa
and the net calorific value of the fuel, Qnet , p  f e. Therefore, for the fuel air ratio for the

system becomes:

m f Qin ,a
fa   … (52)
m a Qnet , p  f

The mass flow rate of fuel ( m f ) can be calculated using equation (52).

Another important parameter which can be used to express the performance of an


actual gas turbine cycle is the Specific Fuel Consumption (SFC), i.e., fuel mass flow
required divided by net work output. According to Saravanamuttoo et al. (2001, p. 71),
it may be defined in terms of the thermal efficiency of the cycle as follows:

1 f
SFC   a in kg/kJ … (53)
th  Qnet , p  f W net

It may be observed that for constant cycle efficiency, an increase in the net calorific
value of the fuel reduces the specific fuel consumption of the gas turbine. This means
that if a better fuel [in terms of net calorific value] is used, lesser quantity of the fuel

e
Net Calorific Value Q  net , p f is the heat released by the complete combustion per unit of fuel
combusted (under ideal conditions) at constant pressure. It excludes the latent heat of the H2O vapour
as this heat cannot be utilized by a gas turbine (Saravanamuttoo et al., 2001, p. 71).

44
has to be used. The SI unit of SFC is ‘kg/J’ but usually electric power is expressed in
‘kWh’. Usually, calorific values of fuels are expressed in ‘kJ/kg’ and SFC is
expressed in ‘kg/kWh’. Hence, the equation may be restructured to obtain SFC in
‘kg/kWh’ as follows:

3600
SFC  in kg/kWh … (54)
th  Qnet , p  f

If the cost of the fuel is given in pence per litre (in this case, the fuel being used is
kerosene, which is liquid) then the unit of Qnet , p  f which will have to be used is

‘kWh/l’. This quantity is dependent on the temperature of the fuel and the source of
the fuel although these differences are usually neglected for the ease of calculation. In
this case, the unit of SFC changes to ‘l/kWh’.

When comparing the SFC for the simple cycle [Graph (8a)] with the SFC for the heat
exchange cycle [Graph (8b)], it can be seen that at T3 = 1500°C and Π = 4, SFC in the
former case is 0.48 l/kWh and in the latter case is 0.21 l/kWh. This shows that that the
addition of a heat exchanger reduces the fuel consumption considerably (by almost
half, in this case) at low pressure ratios. In both cases, an increase in the turbine inlet
temperature reduces the specific fuel consumption.

According to Spiers (1961, p. 270), the net calorific value of kerosene is 1.45 therm
per U.K. Gallon at 15.5°C.

1 therm = 29.3071 kWh (Spiers, 1961, p. 16)


1 U.K. Gallon = 4.54596 l (Spiers, 1961, p. 12)

Hence, the net calorific value of kerosene at 15.5°C in kWh/l will become:

Q 
net , p f  9.3479 kWh/l

45
Due to improvements in fuel technology, this quantity may have increased slightly.
Hence, the author is assuming an approximate value for Qnet , p  f as 10 kWh/l in the

calculations.

Total Fuel Consumption (FT) for the system may be obtained by using equations (50)
and (54):

FT  SFC  ET (l) … (55)

If the cost of the fuel (in £/l) is CF, then the total fuel costs (in £) can be calculated as
follows:

FT  C F
C F ,total  (£) … (56)
100

The cost of fuel required (in l) to produce 1 kWh of electricity using the gas turbine
system can be calculated using the following expression:

C F ,total CF 3600
C F ,kWh    … (57)
PE  AH  E  th  Qnet , p  f

The fuel costs of the simple cycle have been plotted in Graph (9). In this case, it can
be clearly seen that increasing turbine inlet temperatures can reduce fuel costs
considerably although this increases equipment and maintenance costs as high turbine
inlet temperatures reduce durability of the hot-end components of the microturbine (as
discussed earlier). Hence, a cost-benefit balance needs to be made between turbine
inlet temperatures and component material costs. At low pressure ratios, the fuel costs
are virtually unaffected by differences in T3 but at higher pressure ratios, the costs
increase considerably on the reduction of T3.

46
Similar trends can be observed for fuel costs of the recuperated cycle [Graphs (10a-c)]
but at lower pressure ratios, the recuperated cycle has lower fuel costs than the simple
cycle. At T3 = 1500°C and Π = 4, the fuel cost for the simple cycle is about 18 p/kWh;
while for the heat exchange cycle with heat exchanger effectiveness 75%, 85% or
95%, the fuel costs are about 9, 8 and 6 p/kWh, thereby exhibiting a clear economic
advantage (in terms of fuel cost) of the recuperated cycle.

47
4. CONCLUSION

The objective of the project was to perform a constraint analysis of microturbine


engine with an electrical output 250 kWe and to analyse the conditions and parameters
for which a recuperated cycle (used in most models of microturbine generators) is
best suited for microturbine generators of the given size range. It can be inferred from
the study that at low pressure ratios of about 4:1, the recuperated cycle can provide
lower fuel costs than its unrecuperated counterpart. It is desirable to obtain higher
turbine inlet temperatures in order to increase thermal efficiency of the microturbine
system.

Graphs (11a-c) show the ratio of fuel costs of the simple cycle to those of the
recuperated cycle (with a heat exchanger of effectiveness 75%, 85% or 95% added to
the system). It can be seen that for a turbine inlet temperature of 1500°C, the fuel cost
ratio is greater than 1 in all cases, thereby indicating that the recuperated cycle is
better (in terms of fuel cost) than the simple cycle for all pressure ratios in the range
of 1 to 40. Up to a pressure ratio of 5, the recuperated cycle is definitely better (in
terms of fuel cost) than the simple cycle for turbine inlet temperatures as low as
500°C.

Key technical features which need to be improved upon in order to make microturbine
generators more viable and popular are better and cheaper materials for hot-end
components such as the recuperator and the turbine; achievement of better component
efficiencies leading to better overall efficiencies of greater than 40% and reduction of
toxic emissions. Improved gasification technology can lead the microturbine to such a
stage where it can run on virtually any fuel. The use of microturbines in trigeneration
is a future possibility.

48
Though microturbines have only captured a number of niche markets, the scope of
improvement is unlimited as technologies associated with microturbines are
improving rapidly and because mainstream sources of energy are facing extensive
depletion with little or no chance of recovery. The author believes that microturbines
could replace large power plants all over the world and every house could have its
own power generation unit.

If time had permitted, then more intensive research could have been carried out about
relating the microturbine system. An off-design analysis could be carried out for the
various components of the microturbine. Suitable turbine designs could be explored
and ascertained to provide better turbine efficiencies. An intensive study of
engineering materials applicable to microturbines (especially for the hot-end
components) could be accomplished. A more comprehensive economic analysis could
be performed to ascertain capital costs and operation & maintenance costs for the
microturbine system.

49
5. NOMENCLATURE

AH number of hours (annual)


CF fuel cost
cP specific heat at constant pressure
ET total energy
fa fuel-air ratio
FT total fuel consumption
g acceleration due to gravity
h specific enthalpy
m mass flow rate
m mass flow of fluid
∆p pressure loss
p absolute pressure
P power
Q heat transfer rate

Qin heat supplied per unit mass flow


Qnet, p net calorific value at constant pressure
SFC specific fuel consumption
T absolute temperature
u flow velocity
W specific work
z elevation from reference plane
γ ratio of specific heats
ε effectiveness of heat exchanger
η efficiency
Π compression pressure ratio
Π’, Π’’ expansion pressure ratios for actual cycles

50
Suffixes
1 , 2, 3, etc. reference planes
a dry air
b, B combustion chamber
C compressor
E electrical
e exit
f fuel
g combustion gases
HE, hex heat exchange
i inlet
m, M mechanical
opt optimum
red reduced
T turbine, total
th thermal

51
6. REFERENCES

Borbely, A. and Kreider, J.F. eds. (2001), Distributed Generation: The Power
Paradigm for the New Millennium, 1st ed., Boca Raton: CRC Press.

Bullin, A. (2002), An Introduction to Micro-turbine Generators. In: Moore, M.J., ed.,


Micro-turbine Generators, 1st ed., London: Institution of Mechanical Engineering.

Cengel, Y.A. and Turner, R.H. (2005), Fundamentals of Thermal-Fluid Sciences, 2nd
ed., London: Tata McGraw-Hill.

Dambach, R., Hodson, H.P. and Huntsman, I. (2002), Tip-leakage Flow: A


Comparison between Axial and Radial Turbines. In: Moore, M.J., ed., Micro-turbine
Generators, 1st ed., London: Institution of Mechanical Engineering.

Eastop, T.D. and Croft, D.R. (1990), Energy Efficiency for Engineers and
Technologists, 1st ed., Essex: Longman Group.

Energy Nexus Group (2002), Technology Characterization: Microturbines [Online].


Available: http://www.epa.gov/chp/pdf/microturbines.pdf [2005, November 30].

Gorla, R.S.R. and Khan, A.A. (2003), Turbomachinery: Design and Theory, 1st ed.,
New York: Marcel Dekker.

Håll, U. (2002a), Centrifugal Compressors, Lecture Notes to Cohen, Rogers and


Saravanamuttoo, Gas Turbine Theory, Chapter 4 [Online]. Available:
http://www.tfd.chalmers.se/~ulfh/gas_turb_h/crsohbilder/crs4en.pdf, [2006, April 2].

Håll, U. (2002b), Turbines, Lecture Notes to Cohen, Rogers and Saravanamuttoo, Gas
Turbine Theory, Chapter 7 [Online]. Available:
http://www.tfd.chalmers.se/~ulfh/gas_turb_h/crsohbilder/crs7en.pdf [2006, April 2].

52
Hamilton, S.L. (2003), Microturbine Generator Handbook, 1st ed., Oklahoma:
Pennwell Corporation.

Hodge, J. (1955), Cycles and Performance Estimation (Gas Turbine Series Vol. 1),
London: Butterworths Scientific Publications.

Pullen, K.R., Martinez-Botas, R. and Buffard, K. (2002), Design Problems in Micro-


turbine Generators. In: Moore, M.J., ed., Micro-turbine Generators, 1st ed., London:
Institution of Mechanical Engineering.

Saravanamuttoo, H.I.H., Rogers, G.F.C. and Cohen, H. (2001), Gas Turbine Theory,
5th ed., Harlow: Pearson Education.

Shane, T. (2002), Analysis of Micro- and Mini- turbine Competitive and Supply
Markets in Europe. In: Moore, M.J., ed., Micro-turbine Generators, 1st ed., London:
Institution of Mechanical Engineering.

Shepherd, D.G. (1960), Introduction to the Gas Turbine, 2nd ed., London: Constable
and Company.

Spiers, H.M., ed. (1961), Technical Data on Fuel, 6th ed., London: British National
Committee.

Stares, I.J. and Mabbutt, Q.J. (2002), Design Reliability of Micro-turbines. In: Moore,
M.J., ed., Micro-turbine Generators, 1st ed., London: Institution of Mechanical
Engineering.

Thomson, W.R. (1949), The Fundamentals of Gas Turbine Technology, 1st ed.,
London: Power Jets Limited.

53
7. FIGURES

Ideal Simple Brayton Cycle

Ideal Brayton Cycle with Heat Exchange

54
Actual Simple Brayton Cycle

Actual Brayton Cycle with Heat Exchange

55
8. GRAPHS

1 Effect of Π on Thermal Efficiency of Ideal Brayton Cycle


Effect of T3 and Π on Specific Work Output of Ideal Simple Cycle and Cycle with Heat
2
Exchange
3 Effect of T3 and Π on Efficiency of Ideal Brayton Cycle with Heat Exchange
4a Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 70%, ηT = 85%
4b Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 75%, ηT = 85%
4c Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 80%, ηT = 85%
4d Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 85%, ηT = 85%
4e Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 90%, ηT = 85%
4f Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 95%, ηT = 85%
5a Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 85%, ηT = 70%
5b Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 85%, ηT = 75%
5c Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 85%, ηT = 80%
5d Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 85%, ηT = 85%
5e Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 85%, ηT = 90%
5f Variation in Efficiency of Actual Brayton Cycle with T3 and Π at ηC = 85%, ηT = 95%
6a Variation in Efficiency of Heat Exchange Cycle T3 and Π at ε = 70%
6b Variation in Efficiency of Heat Exchange Cycle T3 and Π at ε = 75%
6c Variation in Efficiency of Heat Exchange Cycle T3 and Π at ε = 80%
6d Variation in Efficiency of Heat Exchange Cycle T3 and Π at ε = 85%
6e Variation in Efficiency of Heat Exchange Cycle T3 and Π at ε = 90%
6f Variation in Efficiency of Heat Exchange Cycle T3 and Π at ε = 95%
Variation in Efficiency of Actual Brayton Cycle and Heat Exchange Cycle (ε = 85%)
7
with Π at T3 = 900°C, ηC = ηT = 85%
8a Variation in SFC of Actual Brayton Cycle with T3 and Π at ηC = 85%, ηT = 85%
Variation in SFC of Actual Heat Exchange Cycle with T3 and Π at ηC = ηT = 85% and ε
8b
= 85%
9 Variation in Fuel Cost of Unrecuperated Microturbine with T3 and Π at ηC = ηT = 85%
Variation in Fuel Cost of Recuperated Microturbine with T3 and Π at ηC = ηT = 85%, ε
10a
= 75%
Variation in Fuel Cost of Recuperated Microturbine with T3 and Π at ηC = ηT = 85%, ε
10b
= 85%

56
Variation in Fuel Cost of Recuperated Microturbine with T3 and Π at ηC = ηT = 85%, ε
10c
= 95%
Variation of Fuel Cost Ratio (Simple/ Recuperated) with T3 and Π at ηC = ηT = 85%, ε
11a
= 75%
Variation of Fuel Cost Ratio (Simple/ Recuperated) with T3 and Π at ηC = ηT = 85%, ε
11b
= 85%
Variation of Fuel Cost Ratio (Simple/ Recuperated) with T3 and Π at ηC = ηT = 85%, ε
11c
= 95%

57
Gr. 1 - Effect of Pressure Ratio on Ideal Cycle Thermal Efficiency

70
γ = 1.4

60

50 ηth tends to 100% for all Turbine Inlet Temperatures at Π = ∞


Efficiency η %

40

30

20

10
At Π = 1, ηth = 0% due to absence of compression or expansion work

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

58
Gr. 2 - Effect of Turbine Inlet Temperature and Pressure Ratio on Specific Work Output of
Ideal Simple Cycle and Cycle with Heat Exchange
2.5
T1 =15°C Optimum Specific Work Output
γ = 1.4 T3 = 1500°C

2.0
W/CP.T1
Work Output

1.5
SpecificOutput
Specific Work

1.0

Increasing Turbine Inlet Temperature T3

0.5

T3 = 500°C

0.0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

59
Gr. 3 - Effect of Turbine Inlet Temperature and Pressure Ratio on Efficiency of the Ideal
Brayton Cycle with Heat Exchange
90
T1 =15°C
γ = 1.4
80
ε = 100%

70

60 T3 = 1500°C
Efficiency η %
Efficiency η %

50

40
Increasing Turbine Inlet Temperature T3
30

20

T3 = 500°C
10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

60
Gr. 4a - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 70% ηT = 85%
70
ηC = 70%
ηT = 85%
60

50
Efficiency η %

40
T3 = 1500°C

30
T3 = 500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

61
Gr. 4b - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 75% ηT = 85%
70
ηC = 75%
ηT = 85%
60

50

T3 = 1500°C
Efficiency η %

40

30
T3 = 500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

62
Gr. 4c - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 80% ηT = 85%
70
ηC = 80%
ηT = 85%
60

50

T3 = 1500°C
Efficiency η %

40

30
T3 = 500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

63
Gr. 4d - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 85% ηT = 85%
70
ηC = 85%
ηT = 85%
60

50

T3 = 1500°C
Efficiency η %

40

30
T3 = 500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

64
Gr. 4e - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 90% ηT = 85%
70
ηC = 90%
ηT = 85%
60

50

T3 = 1500°C
Efficiency η %

40

30
T3 = 500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

65
Gr. 4f - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 95% ηT = 85%
70
ηC = 95%
ηT = 85%
60

T3 = 1500°C
50
Efficiency η %

40

30
T3 = 500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

66
Gr. 5a - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 85% ηT = 70%
70
ηC = 85%
ηT = 70%
60

50
Efficiency η %

40

30
T3 = 500°C
T3 = 1500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

67
Gr. 5b - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 85% ηT = 75%
70
ηC = 85%
ηT = 75%
60

50
Efficiency η %

40

T3 = 1500°C
30
T3 = 500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

68
Gr. 5c - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 85% ηT = 80%
70
ηC = 85%
ηT = 80%
60

50

T3 = 1500°C
Efficiency η %

40

30
T3 = 500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

69
Gr. 5d - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 85% ηT = 85%
70
ηC = 85%
ηT = 85%
60

50

T3 = 1500°C
Efficiency η %

40

30
T3 = 500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

70
Gr. 5e - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 85% ηT = 90%
70
ηC = 85%
ηT = 90%
60

T3 = 1500°C
50
Efficiency η %

40

30
T3 = 500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

71
Gr. 5f - Variation in Efficiency of Actual Brayton Cycle with Turbine Inlet Temperature and
Pressure Ratio at ηC = 85% ηT = 95%
70
ηC = 85%
ηT = 95% T3 = 1500°C
60

50
Efficiency η %

40

30
T3 = 500°C

20

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

72
Gr. 6a - Variation of Efficiency of Heat Exchange Cycle with Pressure Ratio and Turbine Inlet
Temperature at ε = 70%
70
ηC = ηT = 85%
ε = 70%
60

50
T3 = 1500°C
Efficiency η %

40

30

20

10
T3 = 500°C

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

73
Gr. 6b - Variation of Efficiency of Heat Exchange Cycle with Pressure Ratio and Turbine Inlet
Temperature at ε = 75%
70
ηC = ηT = 85%
ε = 75%
60

50
T3 = 1500°C
Efficiency η %

40

30

20

10
T3 = 500°C

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

74
Gr. 6c - Variation of Efficiency of Heat Exchange Cycle with Pressure Ratio and Turbine Inlet
Temperature at ε = 80%
70
ηC = ηT = 85%
ε = 80%
60

50
T3 = 1500°C
Efficiency η %

40

30

20

10
T3 = 500°C

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

75
Gr. 6d - Variation of Efficiency of Heat Exchange Cycle with Pressure Ratio and Turbine Inlet
Temperature at ε = 85%
70
ηC = ηT = 85%
ε = 85%
60

50
T3 = 1500°C
Efficiency η %

40

30

20

10
T3 = 500°C

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

76
Gr. 6e - Variation of Efficiency of Heat Exchange Cycle with Pressure Ratio and Turbine Inlet
Temperature at ε = 90%
70
ηC = ηT = 85%
ε = 90%
60

50
T3 = 1500°C
Efficiency η %

40

30

20

10
T3 = 500°C

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

77
Gr. 6f - Variation of Efficiency of Heat Exchange Cycle with Pressure Ratio and Turbine Inlet
Temperature at ε = 95%
70
ηC = ηT = 85%
ε = 95%
60

50
T3 = 1500°C
Efficiency η %

40

30

20

10
T3 = 500°C

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

78
Gr. 7 - Variation in Efficiency of Actual Brayton Cycle and Heat Exchange Cycle (ε = 85%)
with Π at T3 = 900°C, ηC = ηT = 85%

40

Thermal Efficiency of Simple Cycle


30
Efficiency η %

20

Thermal Efficiency of Heat Exchange Cycle

10

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

79
Gr 8a - Variation in SFC of Actual Brayton Cycle with T3 and Π at ηC = 85%, ηT = 85%

1.0

0.8

0.6
SFC l/kWh

0.4

T3 = 500°C

0.2

T3 = 1500°C

0.0
0 5 10 15 20 25 30 35 40

Pressure Ratio Π

80
Gr 8b - Variation in SFC of Actual Heat Exchange Cycle with T3 and Π at ε = ηC = ηT = 85%

1.0

T3 = 500°C
0.8
SFC l/kWh

0.6

0.4

0.2
T3 = 1500°C

0.0
0 5 10 15 20 25 30 35 40

Pressure Ratio Π

81
Gr. 9 - Variation in Fuel Cost of an Unrecuperated Microturbine with T3 and Π at ηC = ηT =
85%
100

80

T3 = 500°C

60
Fuel Cost p/kWh

40

20

T3 = 1500°C
0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

82
Gr. 10a - Variation in Fuel Cost of Recuperated Microtubine with T3 and Π at ε = 75%

100
ηC = ηT = 85%
ε = 75%

80

T3 = 500°C
60
Fuel Cost p/kWh

T3 = 1500°C
40

20

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

83
Gr. 10b - Variation in Fuel Cost of Recuperated Microtubine with T3 and Π at ε = 85%

100
ηC = ηT = 85%
ε = 85%

80

T3 = 500°C
60
Fuel Cost p/kWh

T3 = 1500°C
40

20

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

84
Gr. 10c - Variation in Fuel Cost of Recuperated Microtubine with T3 and Π at ε = 95%

100
ηC = ηT = 85%
ε = 95%

80

T3 = 500°C
60
Fuel Cost p/kWh

T3 = 1500°C
40

20

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

85
Gr. 11a - Variation of Fuel Costs Ratio (Simple/ Recuperated) with Turbine Inlet Temperature
and Pressure Ratio at ε = 75%

5
ηC = ηT = 85%
ε = 75%

4
Fuel Cost Ratio

T3 = 1500°C
2

T3 = 500°C

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

86
Gr. 11b - Variation of Fuel Costs Ratio (Simple/ Recuperated) with Turbine Inlet Temperature
and Pressure Ratio at ε = 85%

5
ηC = ηT = 85%
ε = 85%

4
Fuel Cost Ratio

T3 = 1500°C
2

T3 = 500°C

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

87
Gr. 11c - Variation of Fuel Costs Ratio (Simple/ Recuperated) with Turbine Inlet Temperature
and Pressure Ratio at ε = 95%

5
ηC = ηT = 85%
ε = 95%

4
Fuel Cost Ratio

T3 = 1500°C
2

T3 = 500°C

0
0 5 10 15 20 25 30 35 40
Pressure Ratio Π

88

You might also like