Chapter 5

Conventional Power Generating Systems

Abstract

This chapter provides an overview and up-to-date discussion of conventional power generating systems, through the key cycles, with a prime focus on energy and exergy analyses. Some environmental and economic aspects of their operations are also discussed. Centralized electricity generation stations started to develop in the last quarter of nineteenth century, when the steam turbine and the electrical light bulb had been invented and boosted societal evolution toward the electrification era. The overwhelming majority of power stations in use today are based on the steam Rankine cycle. Everywhere around the world, conventional steam Rankine power plants convert the energy of fossil fuels (mostly coal) and nuclear energy into electricity. Besides steam Rankine cycle there are other vapor power cycles that became important in the last decades, namely organic Rankine cycles; these are introduced briefly in the chapter. Other types of conventional power generation systems are based on gas power cycles. Likewise, the Brayton cycle and its variants, the Otto, Diesel, Stirling, and Ericson cycles, are introduced in this chapter. Nuclear power generating stations are treated in a separate chapter due to their specificity. Thermodynamic cycles are introduced gradually, starting with the most basic ones (e.g., the four-component Rankine cycle) and continuing with improved cycles such as reheat and regenerative cycles. In the last part of the chapter another category of conventional power generation systems is briefly discussed, namely hydropower stations; their energy and exergy analyses are described through illustrative examples and case studies.

Keywords

Energy

Exergy

Efficiency

Exergy destruction

Steam cycles

Organic Rankine cycles

Gas power cycles

Combined cycle power plants

Hydropower

Carbon emissions

Power generation

Nomenclature

BWR back work ratio

EDF exergy destruction fraction

ER expansion ratio

Ex˙si236_e exergy rate, W

ex specific exergy, kJ/kg

FWR flow work ratio

f flow fraction

h specific enthalpy, kJ/kg

m˙si237_e mass flow rate, kg/s

NTU number of thermal units

n˙si238_e molar flow rate, kmol/s

P pressure, kPa

PR pressure ratio

Q˙si239_e heat rate, W

q mass-specific heat flux, kJ/kg

S˙si240_e entropy rate, W/K

SSP specific size parameter, kg/kJ

s mass-specific entropy, kJ/kg.k

T temperature, K

v specific volume, m3/kg

W˙si241_e work rate, W

w mass-specific work, kJ/kg

x vapor quality

Greek letters

γ  specific heat ratio

η  energy efficiency

ψ  exergy efficiency

θ  dimensionless temperature

Subscripts

0 reference state

b boiling

c condensation

C Carnot (factor)

cons consumed

d destroyed

deliv delivered

drop pressure drop

el electrical

gen generated

in input

loss losses

m mechanical

mix mixer

net net generation

l liquid

out output

p pump

ph preheating

R regenerative

rev reversible

rh reheating

sat saturation

sc subcooling

si sink

sh superheating

so source

t turbine

trap steam trap

v vapor

vg vapor generation

Superscripts

* throat condition

¯si242_e average value

5.1 Introduction

Power generating systems are generally treated as heat engines to convert heat input into work, hence to produce electricity at a sustained rate. Heat input is supplied by burning fossil fuels (coal, oil and natural) and biomass, or processing nuclear fuel, or harvesting thermal energy from renewable energy sources. For example, in a conventional coal-fired power plant (the term power station is also used), the energy of coal is eventually converted into power. In general, conventional power stations comprise multiple generating units which are designed to operate at their nominal load when they function optimally.

There are a number of well-known power generating systems denoted as conventional, namely the spark ignition engine, compression-ignition engine, steam Rankine or organic Rankine power plant, combustion turbine power plant, combined cycle power station, nuclear power station, and hydroelectric power station. All these conventional power generating systems (CPGSs) primarily produce mechanical work which is transferred to subsequent systems in the form of shaft rotation. In vehicles, shaft power developed by engines is transferred to the traction system for propulsion. In stationary power plants or generators the shaft power developed by the prime mover is used to rotate an electrical generator which converts the rotational mechanical power to electrical power.

The key component of a CPGS is the prime mover or the organ that produces shaft power. Two types of prime movers are used in CPGSs: positive displacement machines (e.g., reciprocating engines) and turbomachines. Reciprocating machines generally consist of piston-and-cylinder assemblies where the pressure force of an expanding gas is transformed in a reciprocating movement which subsequently is converted into shaft rotation. Turbomachines (turbines) convert kinetic energy of a fluid directly into shaft rotation.

Small-scale CPGS use in general reciprocating prime movers; these are the spark ignition engine and the compression-ignition engine. Large-scale CPGS use turbines as prime movers. The only CPGS which does not use heat as an energy source is the hydroelectric power plant, where hydraulic energy is the input. All other CPGS represent thermomechanical converters and operate based on a specific thermodynamic cycle. The steam Rankine cycle is used in coal-fired, gas-fired, and oil-fired power stations and conventional nuclear power plants. The Brayton cycle is used in gas turbine power plants. A diesel cycle is specific to compression-ignition engines, whereas the spark ignition engine operates based on the Otto cycle.

Any CPGS has its distinct type of equipment. As already mentioned, the most important equipment is the prime mover: steam power plants develop power with the help of steam turbines, gas turbine power plants develop power using specific turbomachinery as the prime mover (this is the gas turbine), hydropower plants use various types of hydraulic turbines, and internal combustion engines use reciprocating piston–cylinder systems for their admission, combustion, compression, and expansion processes, thus generating net work output.

In steam power plants the second major piece of equipment after the steam turbine is the steam generator. Conventional steam generators use to be fired with coal, oil, or natural gas. In a nuclear power plant the steam generator is more specialized, as it is heated using various types of systems aimed at transferring heat from the nuclear reactor to the boiling water in a controlled and safe manner. The specific nuclear-based power generating systems and their power cycles, conventional and advanced, are introduced in Chapter 6 of this book.

In this chapter, the CPGSs are presented in the following order: vapor cycle power plants, gas turbine cycle power plants, gas engines, and hydroelectric power stations. For steam power plants the thermodynamic cycle of steam Rankine type is presented first with various arrangements. Coal-fired power stations with their specific steam generators are then introduced. Organic Rankine cycle (ORC) systems are discussed as a variant of Rankine cycles using an organic working fluid instead of steam. The focus is then shifted to gas turbine cycle power plants with analyses of the air-standard Brayton cycle. The section on internal combustion power generating systems covers information about the Diesel, Otto, Stirling, and Ericson cycles. The last section before chapter’s conclusion discusses hydro power plants. More importantly, the CPGSs and their components are analyzed thermodynamically by writing all balance equations for mass, energy, entropy and exergy, and the performance assessments of these systems and components are carried out by energy and exergy efficiencies as well as other energetic and exergetic performance evaluation criteria.

5.2 Vapor Power Cycles

It is general knowledge that steam power emerged as a technology to generate motive force in eighteenth century. The steam Rankine cycles is the most used type among the power generating thermodynamic cycles for power generation worldwide. Another version of Rankine cycle is the organic Rankine cycle, a technology that has been developed to commercial level in the last few decades and is used now in some specific applications, especially in renewable energy systems.

In a first phase of their development (eighteenth to nineteenth centuries), steam power stations operated based on a non-condensing Rankine cycle using a reciprocating steam engine (also known as steam piston). The motive force generated by the non-condensing steam engine fueled with wood or coal has been extensively used in the nineteenth century in many industries, including textile, mining, and agriculture (for irrigation and crop processing), pumping stations, all sorts of mills, and for transport vehicles such as rail locomotives and ships.

Major progress with steam power generation was marked by the invention of the steam turbine by the end of the 1800s. Since then, the steam turbine has become a prime mover which is responsible today for more than 80% of electrical power generated worldwide. The operation of steam power plants is based on Rankine cycles that use steam as working fluid.

5.2.1 The Simple Rankine Cycle

The most basic vapor power cycle is the simple Rankine cycle. This thermodynamic cycle can be described as an ideal internally reversible cycle or as an actual cycle with four irreversible processes. The ideal cycle can be executed with a machine comprising the following components: an ideal pump that operates isentropically and adiabatically (no heat transfer, only work input), an ideal heat exchanger system which plays the role of vapor generator (with no pressure drops and heat losses), a turbine able to operate isentropically and adiabatically (no heat transfer, only work output), and a condenser with no pressure losses and an infinite heat transfer surface (or equivalently an infinite heat transfer coefficient with zero temperature difference between working fluid and heat sink). The vapor generator consists of a preheating section, boiling section, and vapor superheating section. When the working fluid is steam—which is the case for any conventional type of steam power plant—the system is denoted as “steam Rankine cycle.” Note that when an organic fluid or another type of working fluid is used the allure of the T–s diagram might be quite different, depending on the type of the fluid.

The basic configuration (four-components) of the ideal Rankine cycle and the cycle's T–s diagram are presented in Figure 5.1. The cycle comprises four processes: (i) isentropic pressurization of the working fluid in saturated liquid state, process 1–2, (ii) isobaric vapor generation, process 2–3 (with sub-processes of heating 2–3a, boiling 3a–3b and superheating 3b–3), (iii) isentropic expansion of the superheated working fluid, and (iv) a condensation process of the working fluid until it reaches saturated liquid state, process 4–1.

f05-01-9780123838605
Figure 5.1 Basic steam Rankine power plant representation showing the power plant schematics and the simple, ideal Rankine cycle.

The ideal Rankine cycle is internally reversible but it is not a totally reversible cycle because the heat addition process at the source side is not isothermal. Therefore, the cycle as shown in Figure 5.1 has external irreversibility due to heat transfer at the source side. The heat transfer irreversibilities in this case are determined by the temperature difference T3 − T, where T represents the temperature of the working fluid which increases along the process 2–3a–3b–3. This involves a heat transfer with finite temperature difference between the heat source and working fluid. As the cycle is not externally reversible, an exergy efficiency smaller than 1 must be assigned to the cycle. The thermodynamic analysis can be pursued according to first and second laws.

The mass and energy balance equations for each of the components and the overall balances for the ideal cycle are given in Table 5.1. For the case of the ideal Rankine cycle there are no exergy destructions in condenser, pump, or turbine. However, provided that the heat source is at constant temperature, the externally irreversible Rankine cycle shows exergy destructions in the vapor generator which can be estimated based on the temperature difference between heat source and working fluid.

Table 5.1

Mass and Energy Balance Equations for the Ideal Rankine Cycle

Comp.Balance equationsComp.Balance equations
Pump (1–2)MBE: m˙1=m˙2=m˙si28_e
EBE: W˙p=m˙h2h1si29_e
Vapor generator (2 – 3)Preheating
(2 – 3a)
MBE: m˙2=m˙3a=m˙si30_e
EBE: Q˙ph=m˙3ah3ah2si31_e
Turbine (3–4)MBE: m˙3=m˙4=m˙si32_e
EBE: W˙t=m˙h4h3si33_e
Boiling
(3a-3b)
MBE: m˙3a=m˙3b=m˙si34_e
EBE: Q˙b=m˙h3bh3asi35_e
Condenser (4–1)MBE: m˙4=m˙1=m˙si36_e
EBE: Q˙c=m˙h4h1si37_e
Superheating
(3b – 3)
MBE: m˙3b=m˙3=m˙si38_e
EBE: Q˙sh=m˙h3h3bsi39_e
Overall cycleEBE: W˙p+Q˙sg=Q˙c+W˙tsi40_e, where Q˙sg=Q˙sh+Q˙b+Q˙sh=m˙h3h2si41_e

t0015

Let us denote the temperature of the heat source Tso. When the ideal Rankine cycle is analyzed, the heat source can be considered at a constant temperature which is equal to the highest temperature of the working fluid. The heat source Tso provides heat input to the processes of preheating, boiling and superheating.

In other words, the vapor generator is assumed with an “infinite surface for heat transfer” which assures that the working fluid reaches the heat source temperature level in the thermodynamic state 3 (see the cycle in Figure 5.1); T3 = Tso. On the other hand the condensation temperature ideally must be equal to the sink temperature; T1 = T4 = Tsi.

In addition, the sink temperature is ideally the same as the reference temperature; Tsi = T0. The mass flow rate is the same for all system state points, hence it can be denoted generally with m˙=m˙1=m˙2=si243_e. Furthermore, there is no entropy generation or exergy destruction in the condenser, pump or turbine. For the vapor generator the integrals in EnBE and ExBE can be solved directly because Tso = const. For the overall cycle the net power is:

W˙net=W˙t-W˙p

si244_e  (5.1)

and total heat input is:

Q˙vg=Q˙sh+Q˙b+Q˙sh=m˙(h3-h2)

si245_e  (5.2)

In Chapter 1 three general efficiency definitions that can be applied to Rankine cycles were given. The energy efficiency definition was given in Equation (1.67) and the exergy efficiency definition given either by Equation (1.68) or (1.69), which are equivalent. From Equations (5.1) and (5.2) the energy efficiency of the cycle can be defined in terms of net work output divided by the total heat input at source side (vapor generator), namely

η=W˙netQ˙in=W˙t-W˙pQ˙vg

si246_e  (5.3)

The exergy efficiency in Equation (1.68) expresses the ratio between the net exergy delivered (net work output) and the actual exergy consumed. In the Rankine cycle the exergy consumed is the same as the exergy delivered by the heat source. Therefore, the exergy efficiency is given by

ψ=W˙netE˙xin=W˙t-W˙pE˙xso

si247_e  (5.4)

The exergy efficiency when the heat source is of constant temperature is:

ψ=W˙net1-T0TsoQ˙so=ηηC

si248_e  (5.5)

where ηC is the Carnot factor of the constant temperature heat source, ηC = 1 − T0/Tso.

The specific exergy of the flow represents a thermomechanical exergy calculated based on the specific enthalpy and entropy of the working fluid at a given thermodynamic state. According to its definition (see Chapter 1), the exergy is calculated with

ex=h-h0-T0s-s0

si249_e

Here, index 0 represents the reference state which is assumed to be that of water at T0 = 25 ° C and P0 = 101.325 kPa. The variation of specific exergy Δex for any process is calculated with:

Δex=ΔE˙xm˙=exout-exin

si250_e

where m˙si251_e is a reference mass flow rate and subscripts “out” and “in” refer to outlet and inlet conditions, respectively. For each process within the ideal Rankine cycle either heat or work is exchanged between the working fluid and the surroundings. Processes that exchange work with the surroundings are the pumping and the expansion; the specific work is determined with:

w=W˙m˙=hout-hin

si252_e

Here, w is the mass-specific work. Similarly, the heat exchange processes that occur in the preheater, boiler, superheater, and condenser are described by the following equation, which expresses the mass-specific heat flux:

q=Q˙m˙=hout-hin

si253_e

Note that if heat or work is transferred out of the cycle then the sign is negative, whereas if heat or work is received by the working fluid the amount has a positive sign. Thus the work of the pump is negative, and that of the turbine is positive. Heat transfer at the preheater, boiler, and superheater is positive, whereas at the condenser it is negative.

In thermodynamic state 2 the working fluid must have the same specific entropy as in state 1 and the same pressure as in state 3. In thermodynamic state 3a the liquid is saturated at the specified boiling temperature Tb. In state 3b there is a saturated vapor at temperature Tb. State 3 is superheated vapor with the temperature equal to Tb + ΔTsh. State 4 represents the expanded working fluid at pressure equal to P1, with the same specific entropy as in state 3. The vapor quality x (required for state 4) is related to specific enthalpy at the state point based on the equation

h=1-xhl+xhv

si254_e

where hl and hv are the specific enthalpies of saturated liquid and vapor, respectively, at the pressure corresponding to the thermodynamic state.

The mass, energy, entropy and exergy balance equations for the actual Rankine cycle are given in Table 5.2. Some important parameters are defined in Table 5.3. The back work ratio (BWR) represents the ratio of the work required to turn the pump to the work delivered by the turbine. As observed, in the Rankine cycle the BWR is extremely low; below 1%. This is very favorable for high efficiency of power generation. The work necessary to pressurize liquid (pump work) is negligible with respect to the work generated by expansion of gas (turbine work). In this respect, it is an advantage of the Rankine cycle to operate with a working fluid that changes the phase from subcooled liquid to superheated vapor. Other types of thermodynamic cycles that operate with gas have a BWR higher than 30%.

Table 5.2

Balance Equations for the actual Rankine Cycle of Basic Configuration

ComponentBalance equationsComponentBalance equations
Pump
(1–2)
MBE: m˙1=m˙2si1_e
EBE: m˙1h1+W˙p=m˙2h2si2_e
EnBE: m˙1s1+S˙g=m˙2s2si3_e
ExBE: m˙1ex1+W˙p=m˙2ex2+E˙xdsi4_e
Vapor generator (2–3)Preheating
section (2–3a)
MBE: m˙2=m˙3asi5_e
EBE: m˙2h2+Q˙ph=m˙3ah3asi6_e


EnBE: m˙2s2+23adQ˙phTso+S˙g=m˙3as3asi7_e


ExBE: m˙2ex2+23a1T0TsodQ˙ph=m˙3aex3a+E˙xdsi8_e
Turbine
(3–4)
MBE: m˙3=m˙4si9_e
EBE: m˙3h3=m˙4h4+W˙tsi10_e
EnBE: m˙3s3+S˙g=m˙4s4si11_e
ExBE: m˙3ex3=m˙4ex4+W˙t+E˙xdsi12_e
Boiling
section (3a–3b)
MBE: m˙3a=m˙3bsi13_e
EBE: m˙3ah3a+Q˙b=m˙3bh3bsi14_e

EnBE:m˙3as3a+3a3bdQ˙bTso+S˙g=m˙3bs3bsi15_e


ExBE: m˙3aex3a+3a3b1T0TsodQ˙b=m˙3bex3b+E˙xdsi16_e
Condenser
(4–1)
MBE: m˙4=m˙1si17_e
EBE: m˙4h4=m˙1h1+Q˙csi18_e

EnBE: m˙4s4+S˙g=m˙1s1+41dQ˙cTsisi19_e

ExBE: m˙4ex4=m˙1ex1+411T0TsidQ˙c+E˙xdsi20_e
Superheating
section (3b–3)
MBE: m˙3b=m˙3si21_e
EBE: m˙3bh3b+Q˙sh=m˙3h3si22_e


EnBE: m˙3bs3b+3b3dQ˙shTso+S˙g=m˙3s3si23_e



ExBE: m˙3bex3b+3b31T0TsodQ˙sh=m˙3ex3+E˙xdsi24_e
Overall cycleMBE: there is no MBE because no mass crosses the boundary of the overall system.
EBE:W˙p+Q˙sh+Q˙b+Q˙sh=Q˙c+W˙t;EnBE:23adQ˙phTso+3a3bdQ˙bTso+3b3dQ˙shTso+S˙g=41dQ˙cTcsi25_e
ExBE:W˙p+E˙xso=W˙t+411T0TsidQ˙c+E˙xdsi26_e
whereE˙xso=23a1T0TsodQ˙ph+3a3b1T0TsodQ˙b+3b31T0TsodQ˙shsi27_e

t0010

Note: MBE, mass balance equation; EBE, energy balance equation; EnBE, entropy balance equation; ExBE, exergy balance equation; Tso temperature at heat source [K]; Tsi temperature at heat sink [K].

Table 5.3

Summary of Important Parameters for the Rankine Cycle

QuantityDefinitionUnit
Heat inputQ˙in=Q˙ph+Q˙b+Q˙shsi42_ekW
Net work outputW˙net=W˙tW˙psi43_ekW
Energy efficiency, Equation (1.67)η=W˙netQ˙insi44_e%
Exergy deliveredE˙xdeliv=W˙netsi45_ekW
Exergy efficiency, Equation (1.68)ψ=E˙xdelivE˙xconssi46_e%
Carnot factorηC=1T0T3si47_e%
Exergy efficiency (ideal maximum)ψ=W˙netW˙rev=ηηCsi48_e%
Back work ratioBWR=W˙pW˙tsi49_e%
Pressure ratioPR=P2P1si50_e-
Expansion ratioER=v4v3si51_e-

The pressure ratio (PR) quantifies the ratio between high and low pressure of the Rankine cycle (see Table 5.3). In the steam Rankine cycle the value of PR is quite high due to the fact that steam expands in vacuum. Hence, it is necessary to use a turbine with multiple expansion stages to make the actual expansion process more efficient. Moreover, the parameter ER (expansion ratio) compares the vapor specific volume at turbine exit to the volume at turbine inlet (see Table 5.3). In the Rankine cycle the volume of expanded flow is more than one hundred times larger than the steam volume at the turbine inlet. The expansion ratio is a crucial parameter for turbine design and also quantifies the measure in which the size of low-pressure steam pipes at the turbine outlet must be larger than the high-pressure steam pipes at the turbine inlet.

Example 5.1

We consider in internally reversible Rankine cycle of basic configuration with condensation temperature equal to the reference temperature of Tc = 25 °C and boiling occurring at Tb = 200 °C while the superheating degree is ΔTsh = 200 °C. The cycle will be analysed thermodynamically according to energy and exergy.

Given that this is an ideal Rankine cycle, it is implicit that all internal processes and the heat transfer process at the condenser are reversible. The thermodynamic states 1–4 of the ideal Rankine cycle can be easily determined from the known parameters TcTb, and ΔTsh.

In Table 5.4 the calculation steps for the parameters in each of the thermodynamic states of the cycle are briefly explained. Because the mass flow rate is the same in all state points of the cycle all balance equations can be solved in a mass-specific form; intensive properties are therefore used, such as specific enthalpies, entropies, and volumes.

Table 5.4

Thermodynamic States of Ideal Rankine Cycle for Example 5.1

StateState descriptionExemplary case study parameters for cycle represented in Figure 5.1
P(kPa)T(K)v(m3/kg)xh(kJ/kg)s(J/kg.K)ex(kJ/kg)
1Saturated liquid at Tc3.169298.20.0010030104.80.3670
2Subcooled liquid at s2 = s1 and P2 = P31554298.20.001002− 0.38106.30.3671.555
3aSaturated liquid at Tb1554473.20.0011560852.42.331162.1
3bSaturated vapor at Tb1554473.20.1273127936.431879.7
3Vapor superheated with ΔTsh1554673.20.19581.23832557.2521097
4Two-phase fluid at Tc and s4 = s33.169298.236.450.84121577.2520

t0025

For the ideal steam Rankine cycle it is assumed that in state 1 the liquid at pump intake is in a saturated state. Thus, based on condensation temperature and a vapor quality x1 = 0 all other parameters such as specific volume, specific enthalpy, specific entropy, and specific exergy of the working fluid can be determined with the help of a selected equation of the state for steam. A highly accurate equation of the state for water and steam is presented in Wagner and Pruss (1993); tabular steam data can be also found in many thermodynamic textbooks, such as Cengel and Boles (2010) or Moran et al. (2011). The Engineering Equation Solver software has implemented advanced equations of state for thermophysical and thermochemical properties of many other substances; this software was used to calculate numerical data within this book (see Klein, 2012).

Note that the vapor quality is by definition a quantity ranging from 0 (liquid) to 1 (vapor). However, in Table 5.4 the value of x goes beyond this limit in two cases. It is given a negative value for x at state 2; the negative sign indicates a subcooled liquid when h < hl (x2 = − 0.38; the magnitude of x2 is proportional with the degree of subcooling). As well, a value of x superior to 1 at state 3 indicates a superheated vapor when h > hv (x3 = 1.24; the magnitude of x3 is proportional with the degree of superheating).

The cycle diagram in Ts coordinates is given in Figure 5.2. As seen, when represented at scale, state 2 confuses with state 1 and the preheating line is superimposed over the saturated liquid line; this behavior is common for steam Rankine cycles. The processes specific to the studied cycle are indicated in Table 5.5. The summary of the calculated cycle parameters are given in Table 5.6. It is remarked that exergy efficiency is 62.5%, whereas the energy efficiency is only 34.8%.

f05-02-9780123838605
Figure 5.2 The Ts diagram of internally reversible Rankine cycle studied for Example 5.1.

Table 5.5

Processes of Ideal Rankine Cycle from Example 5.1

ProcessProcess descriptionParameters for cycle case study in Example 5.1
q (kJ/kg)w (kJ/kg)Δex (kJ/kg)
1–2Isentropic pumping s2 = s1P2 = P3wp = h2 − h1N/A1.5551.555
2–3aIsobaric (pre)heating x3a = 0, T3a = Tbqph = h3a − h2746.1N/A160.6
3a–3bIsobaric/isothermal boiling x3b = 1, qb = h3b − h3a1940N/A717.6
3b–3Isobaric superheating P3 = P3bqsh = h3 − h3b462.3N/A217.7
3–4Isentropic expansion s4 = s3P4 = P1wt = h3 − h4N/A10971097
4–1Isobaric/isothermal condensation T4 = T1qc = h4 − h12053N/A0.0

t0030

Table 5.6

Summary of Calculated Cycle Parameters for Example 5.1

QuantityValueUnitQuantityValueUnit
Mass-specific heat input, qin3149kJ/kgExergy efficiency, ψ62.5%
Net specific work output, wnet1096kJ/kgBack work ratio, BWR0.14%
Energy efficiency, η34.8%Pressure ratio, PR490
Carnot factor, ηC55.7%Expansion ratio, ER186

t0035

5.2.2 Exergy Destructions in Rankine Power Plants

It is made clear from the above discussion and Example 5.1 that the Rankine cycle is externally irreversible. This is generally valid for any vapor power cycle which includes sensible heating processes such as liquid preheating and vapor superheating because these heat addition processes are non-isothermal. However, it is theoretically possible to imagine a totally reversible vapor power cycle if the heat addition is for a boiling process only.

In Figure 5.3 a totally reversible vapor power generator that operates based on a Carnot cycle (two isothermal and two adiabatic processes) is exemplified. At state 1 a two-phase water-steam mixture at 350 K having a vapor quality of ~0.2 is enclosed in a cylinder and piston assembly. Due to presence of vapor, this mixture is compressible. The piston is moved upward slowly, performing an adiabatic–isentropic compression process (1–2) during which steam liquefies due to pressurization. In state 2 all vapors liquefies and only liquid is found in the cylinder. To perform the compression process a work input to the cycle is required. Further, during process 2–3 heat is added to the working fluid isothermally.

f05-03-9780123838605
Figure 5.3 Totally reversible engine based on a two-phase cycle with steam. Modified from Cengel and Boles (2010).

The piston moves freely in order to accommodate the larger volume of vapors formed during heat addition. No work is performed to the outside during process 2–3. In state 3 only saturated vapor is enclosed in the cylinder. There follows an isentropic expansion process (3–4) during which some moisture is formed until the liquid–vapor mixture reaches a vapor quality of 0.8 (as given in the particular example from Figure 5.3).

Work is generated during the piston stroke which compensates for the compression work, and thus a net work output is produced by the cycle. During the last process (4–1) heat is removed isothermally from the cycle and vapor condenses, thus the volume of the cylinder has to reduce. Although it appears theoretically realizable, this kind of engine has never been constructed because of some insurmountable technical obstacles. For example, during the isothermal heat addition the volume of the working fluid must increase about one thousand times, which is unpractical. Also compressing and expanding in two phases must be performed extremely slowly in order to approach isentropic processes. In this section the irreversibilities and exergy destructions in actual vapor power plants are analyzed and methods for their quantification are given.

Each component of an actual power plant destroys exergy. The most important component of a Rankine power station is the steam turbine, which has the role of converting the energy carried by hot, high-pressure steam into useful shaft power which turns the electric generator. The other components of the plant—major and minor—have the sole role of working jointly to provide a continuous stream of steam at the turbine inlet. The major components of a Rankine power station besides the turbine are the steam generator, the pumps, and the condensers. Other important components are feedwater heaters and a preheater, superheater, reheater, and economizer. Minor components are many: separators, steam traps, drum, drains, valves, check valves, safety valves, etc. There are also components specific to various kinds of power stations—depending on the fuel source—coal, oil, natural gas, or nuclear.

More realistic analysis of the Rankine cycle is also necessary to determine the practically attainable efficiency of a given power plant. In this respect the exergy destruction due to water pumping, steam expansion, heat transfer, and fluid flow resistance must be accounted for. The common way in which irreversibilities are accounted for in power plant analysis is by the use of a number of parameters, as follows:

 Isentropic efficiency of turbines—that expresses the ratio of actual work to isentropic (i.e., reversible) work developed by the turbine

 Isentropic efficiency of pumps—that expresses the ratio of the isentropic (i.e., reversible) work to the actual work required by pumps

 Mechanical efficiency—which quantifies the difference between net amount of work developed by the working fluid and the mechanical work developed by the turbine shaft

 Electrical efficiency—which expresses the conversion efficiency of the mechanical power of the turbine shaft to the electrical power delivered to the consumer (e.g., electrical grid)

 Temperature difference at heat exchangers at source side—which quantifies the difference between the exergy extracted from the heat source and that delivered to the working fluid

 Heat losses at source-side heat exchangers and hot steam conduits—which induces exergy losses and a reduction of net exergy delivered by the heat source to the working fluid

 Temperature difference at condenser—which quantifies the exergy difference between that rejected by the working fluid and that received by the heat sink; there are two cases:

 if the heat sink is a lake, then a minimal temperature difference of ~ 10 °C must be assumed between the condensation temperature and the sink temperature

 if the heat sink is air and a cooling tower is used to reject heat then a temperature difference of a minimum 5 °C must be considered between air and cooling water and in addition a minimum temperature difference between cooling water and condensation has to be assumed; there must be at least a 15 °C temperature difference between condensation and the wet bulb temperature

 Subcooling degree—which is a technical requirement at pump inlet to prevent cavitation issues and malfunction; subcooling also leads to exergy losses with respect to ideal cycle

 Pressure drop in heat exchangers and conduits—which quantifies the loss of exergy due to the friction process between the fluid steam and walls (pipes, conduits, tubes, heat transfer surfaces) in which the fluid is confined or with which it comes in contact

 High-pressure steam leakage rate—which cannot be prevented because the turbine cannot be perfectly sealed and thus additional energy is required to supply additional fresh water as working fluid

 Air penetration rates at condenser—which cannot be stopped because the condenser operates in a vacuum; this increases the pressure in the condenser and reduces turbine work. Moreover, energy consumption is required to de-aerate the condenser and maintain a reasonable lower pressure

 Energy consumption for auxiliary equipment and power plant house—several pieces of auxiliary equipment, such as (depending on the case) conveyors, fuel injectors, fans, pumps, blowers, electric motors, and also lighting and requirements for the power plant house, will diminish the net power delivered to the consumer by a conventional power generation station

In Table 1.8 the balance equations and energy and exergy efficiency formulations for all important devices used in the power generation system are introduced. In this section the exergy analysis presented in Chapter 1 will be applied in order to exemplify how the irreversibilities in steam Rankine cycles can be assessed.

There are at least four additional processes that must be represented on an actual basic Rankine plant diagram that are not present on an ideal plant diagram: water subcooling at pump inlet, heat loss at hot steam conduits, high-pressure steam leak at the turbine, and fresh water addition at the condenser.

These processes are represented in Figure 5.4, which illustrates an actual steam Rankine cycle of basic configuration. The T¯si255_e notations in the figure represent the average system boundary temperature at locations where heat transfer processes occur. In subsequent paragraphs the thermodynamic analysis of the Rankine cycle with irreversibilities is presented and exergy destructions and entropy generations quantified. There are entropy transfer fluxes between surroundings and power plant which can be written mathematically as follows (see Figure 5.4):

S˙ph=+Q˙phTph;S˙b=+Q˙bTb;S˙sh=+Q˙shTsh;S˙loss=-Q˙lossTloss;S˙c=-Q˙cTc;S˙sc=-Q˙scTsc

si256_e

where the sign indicates that entropy flux enters (+) or exits (−) the thermodynamic cycle.

f05-04-9780123838605
Figure 5.4 Representation of a Rankine power plant with multiple exergy destruction processes (finite difference heat transfer, heat losses, pressure drop, electromechanical losses, steam leak).

The thermodynamic analysis of the cycle can start from thermodynamic state 1 where the working fluid is saturated. Three parameters must be specified in order to determine thermodynamic state 1; these are:

 The reference temperature T0 (which can be the temperature of surrounding atmosphere if the condenser is cooled with air, or the lake temperature if the condenser is cooled with water; T0 should be the lowest temperature of power plant surroundings)

 The assumed minimum temperature difference at the heat sink (condenser), ΔTc

 The required subcooling degree ΔTsc necessary for the pump to operate without cavitation.

Thermodynamic state 1 is completely determined by specifying that the liquid is saturated (x1 = 0) and that the temperature of working fluid is given by

T1=T0+ΔTsc+ΔTc

si257_e

The degree of subcooling in practical applications is taken as ΔTsc = 1 − 3 °C, whereas ΔTc = 10 − 15 °C. The condensation pressure is the saturation pressure corresponding to T1. There is negligible pressure drop in any practical subcooling process. However, some pressure drop must be accounted for across the condensation process. Hence, one can take P2 = P1. Further, the temperature of state 2 (subcooled liquid) is determined

T2=T1-ΔTsc

si258_e

State 2 is completely specified because its pressure and temperature can be determined. Thermodynamic state 3 can be determined once the isentropic efficiency of the pump (ηs,p) and the pressure P3 are specified. The isentropic efficiency allows for determination of specific flow enthalpy and is defined by the following equation (see also Chapter 1):

hP3s2-h1=ηs,ph2-h1

si259_e

The type of pump must be chosen in accordance with the head and the flow rate and economic criteria. For low head, centrifugal pumps of multistage configuration are generally preferred. For high head the preferred choice is of double-casing barrel pumps which are also tolerant to high water temperatures. The acceptable range of isentropic efficiency for pumps at power generating stations is 0.8–0.9.

The pressure in state 3 should be superior to the pressure is state 4 (boiling pressure) due to the existence of pressure drop in preheater. In general the preheater is considered as part of the vapor generator, and, depending on the case, it generally consists of an assembly of pipes heated from the exterior and working fluid circulating internally. In some typical cases (e.g., nuclear power stations) the preheater may be just a part of a shell and tube steam generator which has the hot heat transfer fluid circulating in tubes and water being preheated and boiled in the shell side. In typical power plants the pressure drop over preheaters leads to a decrease in saturation temperature of 0.1–0.3 °C. Denoting ΔTsat,ph the decrease of saturation temperature due to pressure drop across preheaters, the pressure in state 3 results from the equation

P3=PsatTb+ΔTsat,ph

si260_e

The pressure in state 4 corresponds to saturation at a specified boiling temperature that typically is in the range of Tb = 100 − 330 °C, depending on the actual type of power generating station. In state 4 there is a saturated liquid, from which all state parameters can be determined based on temperature T4 = Tb and vapor quality x4 = 0. There are many types of boilers, depending on the nature of the heat source. In a coal-fired power plant vertical tubular boilers are used, with boiling occurring in pipes heated by radiation and convection by combustion gases. In oil-fired plants the boilers may be of horizontal return tubular type, again with the boiling process occurring in tubes. Gas-heated power plants have boilers comprising a number of bare or extended-surface tubes forming a heat exchanger register and disposed in a horizontal configuration. In a nuclear power plant the steam generators a pool boiling process is the preferred one.

In any type of boiler there is an unavoidable pressure drop due to working fluid flow. The decrease in saturation temperature due to pressure drop over the boiler (ΔTsat,?b) is on the order of 0.5–1.5 °C, depending on its actual type. Thus the thermodynamic state 5 at boiler exit is fully determined based on its temperature T5 = Tb − ΔTsat,?b and the vapor quality x5 = 1. The decrease in saturation temperature caused by pressure drop during the superheating process which occurs in superheater or reheater (ΔTsat,?sh) is in a practical range of 0.3–0.7 °C. In fossil fuel power stations the superheaters are of tubular type (with or without extended surface), with steam circulated in tubes and heat provided from the exterior. In a nuclear power plant, the superheater may be the same unit as the steam generator. Using an assumed ΔTsat,?sh, the thermodynamic state 6 (Figure 5.4) is determined based on pressure P6 given by

P6=Psat(T5-ΔTsat,sh)

si261_e

and on temperature T6 = Tsh.

The pressure drop in steam conduits of power plants is in the range of ΔPdrop = 0.2 − 0.4 MPa. This pressure drop is due to the high complexity of the steam circuits, which include many types of fittings (screwed, flanged, compensators), headers and manifolds, valves, and check valves.

The steam conduits must be well insulated to minimize the heat losses. It is impossible to avoid heat losses at high-temperature steam conduits because there is generally a considerable distance between the steam generator and the turbine location. In addition, conventional power generating stations have massive steam conduits that must be supported and anchored at many points using special “heat transfer bridges” where heat losses are more accentuated. The temperature reduction of steam due to heat losses at hot steam conduits in power plants can be taken on the order of ΔTloss = 10 − 30 °C. Hence the thermodynamic state 7 at the turbine inlet is determined based on its pressure and temperature.

The turbine’s irreversibilities are accounted for by specifying the exergy efficiency (ηs,?T). Hence, the specific enthalpy of steam at the turbine outlet results from:

h7-h8=ηs,Th7-hP8s7

si262_e

The isentropic efficiency of the turbine has the highest influence on overall cycle efficiency, hence the correct choice of turbine is crucial for a power plant. After Parson’s invention of the reaction steam turbine—in the last quarter of the nineteenth century—steam turbine technology developed fast in the twentieth century. In the first quarter of the nineteenth century there were many types of proprietary steam turbine designs, such as Parson, Westinghouse-Parson, De Laval, Terry, Kerr, Curtis, Hamilton-Holtzworth, and Allis Chalmers. Steam turbines are generally considered highly reliable equipment in conventional power generation stations. They can be classified as impulse and reaction turbines or as condensing, extraction, reheat, and back pressure (non-condensing) turbines. Reheat and condensing turbines are those which are normally used in conventional Rankine power stations. Modern turbines are conceived as a succession of stages of impulse and reaction blades, with impulse-type of stages at higher pressures and reaction-type stages at lower pressures. The isentropic efficiency of modern steam turbines for conventional power stations spans from 0.8 to 0.9.

The pressure at the turbine outlet (state 8) is practically the same as pressure at the condenser inlet. The number of stages of the turbine can be approximated based on the isentropic nozzle formula that correlates the pressure in the throat (P*) with pressure at the nozzle inlet and the specific heat ratio of the steam, γ. For superheated steam γ = 1.14, while saturated steam is generally approximated with γ = 1.3. The general formula for throat pressure is

P*=2γ+1γγ-1Pin

si263_e

from which, using the values of γ, it results that

P*=0.546Pin,forsuperheatedsteam0.576Pin,forsaturatedsteam

si264_e  (5.6)

where Pin stands for pressure at the turbine stage inlet.

Moisture does not actually form inside the steam turbine if the vapor quality is higher than 96% because of the character of the flow, which is a fast flow. Hence, the thermodynamic states in the turbine are metastable: there is not sufficient time for heat transfer or for steam condensation to occur. The thermodynamic locus connecting states with steam quality of 96% (or 4% moisture) is known as the Wilson line. If the thermodynamic state in a turbine stage is above the Wilson line, then the steam is dry. If the thermodynamic state is below Wilson line, in that particular stage steam condensation occurs.

The magnitude of pressure drop on the low-pressure steam line and condenser altogether is of 1–2 kPa, which leads to a saturation temperature drop of ΔTsat, c = 2 − 4 °C. The steam condenser is a massive shell-and-tube heat exchanger with steam condensing at the shell side and with water circulating in pipes.

The mechanical efficiency of the power plant is a parameter that accounts for losses caused by turbine friction, steam leak, and any other mechanical losses at the turbine and is defined by

ηm=wmwnet=0.88-0.94

si265_e

In power generation stations the turbines are optimized to rotate synchronously with the power grid—at 3600 RPM in North America, Japan, and South Korea, where the electrical grid is at 60 Hz and at 3000 RPM and where the grid frequency is 50 Hz. The generator, which is connected in line with the turbine shaft is of synchronous type with a copper-coiled rotor and stator. Both the rotor and the stator generate an important amount of heat. For the stator, heat is removed by cooling from the exterior, but for the rotor it is difficult to reject the generated heat because of the small free space between stator and rotor. The current technology for rotor cooling uses hydrogen as a heat transfer fluid. Hence, an enclosure is built around rotor. The generator is linked to a transformer that raises the electric voltage up to over 100 kW in compatibility with the electrical grid. The electrical efficiency, representing the ratio between electrical energy delivered by the generator–transformer assembly and the mechanical energy transferred by the shaft to the generator is defined by

ηel=welwm=0.9-0.96

si266_e

There are many auxiliary devices in power plants that consume power, and hence diminish, even if in small measure, the net generation. Here are listed some relatively important auxiliary power-consuming processes and devices of conventional power generation stations:

 If condenser is cooled with water from a lake, there must be installed water circulation pumps and water purification systems which require electrical motors to drive pumps

 If condenser is cooled with hyperboloid cooling tower there must be installed water recirculation pumps and water-feeding pumps driven by electrical motors

 If condenser is cooled by mechanical-draft cooling tower, besides the use of electrically driven pumps there must be installed electrically driven fans for air circulation in forced convection

 In coal-fired power stations there are coal conveyers, ash conveyers, coal mills, and pulverizers that consume power

 In natural gas and oil-fired power station the burners must have blowers driven by electrical motors

 Chimneys when installed require electrically driven fans and blowers

 In fossil fuel plants rotary wheel economizers turned by electric motors are installed at exhaust gases

 Some boilers require additional feed pumps

 Turbine auxiliary parts, such as lubrication system and generator cooling system, require some power consumption

 The power station house requires power for lighting, appliances, alarms, and data acquisition, etc.

For a conventional power generating station, the power consumption for all auxiliary devices and processes, including that for the steam power station house, represents a small fraction of the generated electric power, usually 0.5–1%. Other losses such as those caused by steam leakage and air penetration are small and can be neglected with respect to other more important energy losses.

In Table 5.7 the main parameters which quantify the irreversibilities of steam power stations are summarized. The turbine’s isentropic efficiency influences the cycle efficiency to a very large measure. A change of isentropic efficiency of 10% leads to a change of cycle energy efficiency of ~ 3–4% while exergy efficiency change is 8–9%. Temperature difference between the working fluid and the heat source affects the exergy efficiency. As an order of magnitude, a change of temperature of 50 °C may induce a reverse change of exergy efficiency of 1–2%. Heat losses at hot steam conduit inlets influence the cycle’s energy efficiency, but in very small measure, ~ 0.2%. All other irreversibilities have even smaller influence (independently) on energy efficiency of the cycle.

Table 5.7

Parameters That Quantify the Irreversibility of the Rankine Cycle

ParameterName and definitionaTypical range
ηs,pIsentropic efficiency of pump
h3s − h2 = ηs,?p(h3 − h2)
0.8–0.9
ηs,tIsentropic efficiency of turbine
h7 − h8 = ηs,?t(h7 − h8s)
0.8–0.9
ηmMechanical efficiency: ηmwnet = wm0.88–0.94
ηelElectrical efficiency: ηelwm = wel0.9–0.96
ΔTcRequired temperature difference between condensation and heat sink10–15 °C
ΔTscRequired subcooling degree at pumps inlet
ΔTsc = T1 − T2
1–3 °C
ΔTsat, phSaturation temperature decrease due to pressure drop across preheater
ΔTsat,?ph = Tsat,?3 − Tsat,?4
0.1–0.3 °C
ΔTsat, bSaturation temperature decrease due to pressure drop across boiler
ΔTsat,?b = T4 − T5
0.5–1.5 °C
ΔTsat, shSaturation temperature decrease due to pressure drop across superheater/reheater
ΔTsat,?sh = T5 − Tsat,?6
0.3–0.7
ΔPdropPressure drop in hot steam conduits before the turbine
ΔPdrop = P6 − P7
0.2–0.4 MPa
ΔTlossTemperature drop over hot steam conduits due to heat loses
ΔTloss = T6 − T7
10–30 °C
ΔTsat, cSaturation temperature decrease due to pressure drop at sink side
ΔTsat,?c = T8 − T1
2–4 °C

a See Figure 5.4 for notations.

Certain inequality-type relations must be satisfied between state parameters in order comply with the second law of thermodynamics. More specifically, the energy and entropy balance equations for the subcooler (1–2) require

h1=h2+qscands1+sgen=s2+qscT¯sc

si267_e

As the generated entropy must be positive, from the above equations it results that the following inequality must be satisfied

T¯sc<h1-h2s1-s2

si268_e

which reflects the fact that the cooling medium of the subcooler must have an average temperature T¯scsi269_e below a certain value; in addition it is known that T¯sc>T0si270_e.

The energy and entropy balance on the preheater are

h3+qph=h4ands3+qphT¯ph+sgen=s4

si271_e

From the above equation it results that the following inequality must be satisfied

T¯sh>h4-h3s4-s3

si272_e

which states that the average temperature of the heat source temperature at the preheater must be higher than a threshold value that insures that the generated entropy is positive.

Similarly inequalities are derived from the energy and entropy balances on the boiler and superheater as follows:

T¯b>h5-h4s5-s4andT¯sc>h6-h5s6-s5

si273_e

Furthermore, because the preheater, boiler, and superheater are all heated from the same heat source, it results that the following inequalities must be satisfied

T¯sh>T¯b>T¯ph

si274_e

The irreversibilities due to heat losses and pressure drop which occurs at conduits that transport hot, high-pressure steam to the turbine are accounted for by specifying the parameters ΔPdrop and ΔTloss defined in Table 5.7. The energy and entropy balance equations for process 6–7 (see Figure 5.4), which models these types of losses, require that

T¯loss<h6-h7s6-s7

si275_e

The energy and entropy balances for the condenser can be combined to eventually derive the following inequality that must be satisfied

T¯sc<T¯c<h8-h1s8-s1

si276_e

As a conclusion of this section it is clearly noted that the effect of irreversibilities is very important for the steam Rankine cycle and must be quantified accurately in order to predict the actual plant performance. Furthermore, exergy efficiency reveals that the organ which destroys most of the steam exergy is the turbine. Hence, the selection of turbine and its design improvement are crucial issues when designing steam power plants.

The increase of cycle efficiency is very important in practice because it leads to important savings of thermal energy supply (or fuels) and reduction of environmental impact. The engineering goal is to “impinge” the design toward the limits imposed by the second law of thermodynamics. In other words, the design engineer must find ways to reduce the cycle irreversibilities.

In many engineering thermodynamics textbooks some methods to increase the efficiency of the ideal Rankine cycle are discussed. Basic thermodynamic analysis shows that increase of source temperature and/or decrease of sink temperature leads to higher cycle efficiency of any thermodynamic cycle. This is confirmed by a simple analysis of the Carnot factor 1 − T0/T with a decreasing T0 and increasing T. Let us assume that T0 is replaced by T0 − ΔT and T is replaced with T + ΔT. Figure 5.5 reports Carnot factor η when ΔT is varied from 0 to 50 K. For the reference case (ΔT = 0) it is taken that T0 = 25 ° C and T = 200 ° C, in which case ηref = 0.557. Three types of Carnot factors are reported, corresponding to three cases:

f05-05-9780123838605
Figure 5.5 Improvement of Carnot factor by changing source or/and sink temperatures.

 Sink temperature (condensation) is decreased with ΔT, hence η = 1 − (T0 − ΔT)/T.

 Source temperature is increased with ΔT, hence η = 1 − T0/(T + ΔT).

 Both source and sink temperatures are changed, hence η = 1 − (T0 − ΔT)/(T + ΔT).

The results show that it is relatively more important to reduce the sink temperature than to increase the source temperature. If sink temperature decreases by 50 K then efficiency increases by ~ 8%, from 55.7–63.1%. If source temperature is increased with 50 K then the Carnot factor will increase only by 3%. If both temperatures are changed with 50 K then the Carnot factor increases by 10%.

The same trend is observed in the ideal Rankine cycle. In the steam Rankine cycle the condensation temperature is limited by the freezing point of water when working fluid freezes; however, for practical reasons one cannot run water as a working fluid at temperatures lower than 4 − 5 ° C (note that at the bottom of deep lakes and with seawater there is a steady temperature of ~ 4 °C. A simple parametric study is reported in Figure 5.6 which shows the variation of energy efficiency of an ideal steam Rankine cycle when the condensation temperature is decreased by ΔT = 0 … 50 K with respect to a reference value of 55 ° C, and when boiling temperature is increased with same ΔT with respect to a reference value of 175 ° C, and when the temperature of superheated vapor increases from the reference value of 400 ° C with by the amount ΔT; only one parameter is changed at a time while all other cycle parameters remain unchanged.

f05-06-9780123838605
Figure 5.6 Influence of condensation and boiling temperatures and of the degree of superheating on ideal steam Rankine cycle efficiency.

As seen, the reference cycle efficiency is 27%. If condensation temperature is decreased from 55 to 5 °C the cycle efficiency increases to 36.2% (or a 9.2% increase). If boiling temperature increases from 175 to 225 °C the cycle efficiency reaches 31.9% (that is, a 4.9% increase). Note that in the last case the pressure of the boiler increases from 890 to 2550 kPa, while the highest temperature of the cycle is assumed to be constant, T5 = 400 ° C. The increase of boiler pressure due to increase of boiling temperature is illustrated in Figure 5.6. If the boiling temperature is 225 °C and condensation is at 5 °C then the cycle efficiency becomes 40% (or a 13% increase with respect to the reference case). The last curve plotted on the graph in Figure 5.6 shows the variation of energy efficiency when the temperature of superheated vapor, T5, increases from 400 to 450 °C. In this case the cycle efficiency increases relatively less, from 27% to 27.9%.

This simple study suggests a strategy of efficiency improvement for the steam Rankine cycle: first one must set the condensation temperature at the lowest value possible, then one needs to increase the boiling temperature (and pressure) to the highest technical-economic value that is advantageous, and, as a last measure, one must set the temperature of superheated vapor as high as possible based on the practical temperature level that can obtained from the specific heat source considered in the application. Note also that the boiler pressure is limited by technical–economical considerations related to materials selection. For example, in conventional nuclear power plant the boiling pressure is ~5 MPa and the superheated vapor temperature is limited to ~ 350 °C for safety reasons. Furthermore, in a conventional Rankine cycle—which includes a boiling process—the highest pressure should be lower than the critical pressure of steam, which is 22.06 MPa. Newly emerging non-conventional Rankine cycles can operate at supercritical pressure and hence have better efficiency than the conventional cycles.

Example 5.2

We give an example of the thermodynamic analysis of a Rankine cycle with exergy destructions. The cycle configuration and state points are as shown in Figure 5.4. The assumed parameters for the cycle are given in Table 5.8. It is required to determine the entropy generated and exergy destroyed for each process, the exergy destroyed by the cycle, the exergy wasted, the cycle efficiencies, the BWR and expansion ratio, and the number of turbine stages.

Table 5.8

Performance Parameters of a Non-Ideal Rankine Cycle of Basic Configuration

Assumed parametersCycle performance parameters
ParameterValueIrreversibilities
T025 °CComponentsgen (J/kg.K)exd (kJ/kg)(%)
Tb250 °CSubcooler (1–2)0.1930.060.0
Tsh400 °CPump (2–3)2.3110.6890.2
ηs,T0.85Preheater (3–4)222.866.417.3
ηs,P0.85Boiler (4–5)276.082.321.5
ΔTc7 °CSuperheater (5–6)71.021.25.5
ΔTsc2 °CHeat lossesa (6–7)70.020.65.4
ΔTsat,?ph0.2 °CTurbine (7–8)529.7157.941.2
ΔTsat,?b1 °CCondenser (8–1)113.933.98.9
ΔTsat,?sh0.7 °CSpecific energy and exergy flows (kJ/kg)
ΔPdrop0.3 MPaqin = qph + qb + qsh3077
ΔTloss20 °Cwnet = wt − wp926.3
ΔTsat, c3 °Cqwasteqin − wnet2151
Tsc304 Kexin=qph1T0Tph+qb1T0Tsource,b+qsh1T0Tshsi52_e1362
Tph500 Kexd=iexd,isi53_e383.1
Tsource. b550 Kexwaste = exd,surr = exin − wnet − exd52.8
Tsh600 KCycle efficiencies, BWR and ER
Tloss338 Kη = wnet/qin = 30.1%ηC=1T0/T-sh=54.1%si54_eBWR = 0.5%
Tsink,?c305Kψ = η/ηC = 55.6%ψ = wnet/exin = 68%ER = 267

t0045

a Heat losses from hot steam conduits (losses due to pressure drop are included).

The mass balance equations require that the mass flow rate is the same at each state point, m˙i=m˙si277_e. The energy balance equations can be written for each process as indicated in Example 5.1. From energy balances the temperature, pressure, specific volume, specific entropy, and specific enthalpy at each state point can be calculated. From these the specific exergy at each state point can be determined. It is reasonable to assume that the reference state is liquid water at T0 = 25 °C, P0 = 101.325 kPa. Entropy and exergy balance equations give the entropy generated and exergy destroyed for each component, respectively. In Table 5.8 the relevant equations are given together with the results.

It is remarked that the maximum exergy destruction is at the turbine at 41.2%, which is assumed to operate with an isentropic efficiency of 85% per stages (see Table 5.8). The number of stages of the turbine is determined based on Equation (5.6). As seen in the cycle representation in Figure 5.7 the expansion process in the turbine proceeds in seven stages. The first four stages operate above the Wilson line, hence condensation does not occur and impulse blades can be used. The subsequent three stages operate under condensation, when the flow volume increases more prominently across the stage. For this reason the turbine stages must be of reaction type (with expansion both in stator and rotor). The exergy destruction due to internal irreversibilities is denoted with exd in the table. The waste exery represents also a form of exergy destruction, namely, exergy destruction due to system interaction with the surroundings, exwaste = exd,surr.

f05-07-9780123838605
Figure 5.7 Non-ideal steam Rankine cycle of basic configuration (Example 5.2).

Furthermore, the energy efficiency of the non-ideal cycle is as low as 30.1% and the exergy efficiency with respect to the maximum reversible work output (Carnot factor of the heat source) is ψ = η/ηC = 55.6%. The exergy efficiency relative to the actual exergy input into the cycle is ψ = wnet/exin = 68%. If the cycle will operate reversibly under the same conditions, then the energy efficiency is 38.5% and exergy efficiency is 87.2%.

5.2.3 Ideal Reheat Rankine Cycle

Reheating is a method of improving Rankine cycle efficiency which consists of inter-stage heating of the expanding steam. After the first stage of expansion which typically reduces initial steam pressure by one-fourth, the steam is heated up (or close) to the maximum heat source temperature. After the second expansion stage the steam reaches condensation pressure. Although more heat input is required for reheating, the efficiency of a reheating cycle is higher because the two-stage turbine develops relatively more work.

Double-reheating cycles do exist and have been used in conventional power generation plants since the 1950s. With respect to the basic Rankine cycle, the efficiency improvement of using a single reheat is on the order of 5% and using double reheat it is 8%. Hence, using more than two reheat stages does not appear economically justifiable for conventional power generation systems (Cengel and Boles, 2010). The use of a reheating cycle is also justified because in practical applications the maximum temperature of the cycle is constrained for one of two reasons: (i) the nature of the heat source limits the temperature at which heat is available (e.g., geothermal heat, waste heat recovered, concentrated solar receiver with associated temperature of 200–800 °C) or (ii) the safety and material resistance issues (which limits boiler temperature to a maximum of 800 °C for conventional fossil-fuel-based power generation plants and 350 °C for conventional nuclear power plants).

In Figure 5.8 the diagram of a single-reheat Rankine power plant is presented. As seen, after the high-pressure turbine (HPT) the expanded steam is reheated in a special heat exchanger and then further expanded in the low-pressure turbine (LPT). The thermodynamic cycle for the ideal single-reheat Rankine cycle in a Ts diagram is presented in Figure 5.9.

f05-08-9780123838605
Figure 5.8 Single-reheating Rankine power plant.
f05-09-9780123838605
Figure 5.9 Ideal single-reheat Rankine cycle.

One of the features which is important for reheat cycle design is that there is an optimum pressure for the reheating process at which the efficiency of the cycle is maximized for any fixed boiling temperature. This optimum can be observed if the cycle efficiency is plotted against the pressure ratio across the HPT. Figure 5.10 illustrates the existence of the optimum pressure ratio for a relevant case study. Assume that the working fluid is steam and that the highest temperature of the working fluid is fixed to T5 = T7 = 400 °C. The variation of cycle efficiency with pressure ratio PR = P5/P6 is shown for five values of boiling temperature (see Figure 5.10). The optimum pressure ratio is in the range of 3 to 4.6. The vapor quality at the turbine exit decreases for optimum expansion from x8 = 0.94 when Tb = 200 °C to x8 = 0.83 and boiling temperature is 280 °C.

f05-10-9780123838605
Figure 5.10 Variation of energy efficiency of single-reheat cycle with pressure ratio (PR = P5/P6) and boiling temperature (Tb) for fixed superheated and condensation temperatures (Tsh = 400 °C, Tc = 30 °C).

Example 5.3

This example considers a single-reheat ideal Rankine cycle with a configuration similar to that indicated in Figure 5.7. The condensation temperature is Tc = 30 °C, boiling is at 280 °C, and steam is superheated up to 400 °C. Determine the optimum pressure ratio for the HPT, the maximum cycle efficiency, and the mass-specific work for the HPT and LPT and show the Ts diagram of the cycle.

The mass and energy balances for each component of the cycle are written as indicated in Table 5.1. The mass flow rate is the same for all state points, hence the mass-specific quantities can be used. The pressure ratio over the HPT is varied in the range of 3 to 4 and a maximum efficiency of the cycle in the given conditions is determined to be 40.44% for PRopt = 3.066. For this case the mass-specific work developed by the HPT is 280.1 kJ/kg, and the LPT develops 1099 kJ/kg. The cycle diagram is shown in Figure 5.11.

f05-11-9780123838605
Figure 5.11 Ideal single-reheat Rankine cycle for Tc = 30 °C, Tb = 280 °C, Tsh = 400 °C.

5.2.4 Ideal Regenerative Rankine Cycle

At the beginning of the twentieth century, the regenerative Rankine cycle started to be used for power generation with improved efficiency. This cycle is a modification of the basic Rankine cycle. It consists of applying internal regeneration within the cycle aimed at preheating the working fluid after the liquid pressurization process. Regeneration increases the average boiler temperature and makes the ideal Rankine cycle approach better than the Carnot cycle.

Let us compare the ideal Rankine cycle represented in Figure 5.1 with the Carnot cycle. As observed, the process path 3–4–1–2 resembles a part of the Carnot cycle. Also the boiling process 3a–3b is an isothermal heat input such as that in the Carnot cycle. However, compared to the Carnot processes, processes 2–3a and 3b–3 show large abatement because these are non-isothermal heat additions. In particular, process 2–3a has a negative effect on cycle efficiency because it dramatically reduces the temperature at which heat input occurs. The idea of regeneration consists of transferring heat internally from expansion process 3–4 to preheating process 2–3a such that the need for heat input at a lower temperature (below boiling) is reduced or eliminated.

The schematic of the ideal regenerative Rankine cycle is presented in Figure 5.12a. Imagine that the turbine is “surrounded” by a stream of working fluid in liquid phase which receives heat by heat transfer from the expanding flow to increase its temperature according to process 2–3. The rate of heat is indicated in the figure with Q˙phsi278_e. Because there is a heat transfer process superimposed over the isentropic expansion process in the ideal regenerative Rankine cycle the turbine will operate non-adiabatically (and isentropically, that is with no irreversibilities due to expansion) according to a process represented by the line 5–7 in Figure 5.12b.

f05-12-9780123838605
Figure 5.12 Ideal regenerative Rankine cycle with non-adiabatic turbine—(a) power plant schematics, (b) Ts diagram (represented at scale for Tb = 250 ° C, Tc = 30 ° C, ΔTsh = 150 ° C).

The energy balance equation for the regeneration process (see Figure 5.12) can be written with mass-specific quantities as shown below:

qph+wt=h5-h7andqph=h3-h2

si279_e

One also imposes that the ideal regenerative Rankine cycle develops the same turbine work as the internally reversible Rankine cycle of basic configuration operating under the same condition (wt = h5 − h6); hence one has:

h3-h2+h5-h6=h5-h7h6-h7=h3-h2

si280_e

The efficiency of the basic, internally reversible Rankine cycle operating under the same conditions as the ideal regenerative cycle (Tc, Tb, T5 are the same) is expressed as follows:

η=wt-wpqin=h5-h6-h2-h1h5-h2=1-h6-h1h5-h2

si281_e

The efficiency of the ideal regenerative Rankine cycle can be expressed as follows:

ηR=wt-wpqin=h5-h7-h3-h2-h2-h1h5-h3=1-h7-h1h5-h3

si282_e

Comparing η and ηR expressed with the help of above three equations, one has that

ηR=1-h6-h1h5-h3+h3-h2h5-h3>1-h6-h1h5-h2+h3-h2h5-h3>η

si283_e

where one notices that h5 − h2 > h5 − h3 > 0 (see Figure 5.12). Thus, it is demonstrated that the efficiency of the ideal regenerative Rankine cycle is necessarily higher than that of the ideal Rankine cycle of basic configuration operating between the same temperature limits. This increase in efficiency can be as much as 18–20% when the working fluid is steam and the boiling temperature in the range of 200–300 °C.

According to Figure 5.12a, the regenerative Rankine cycle cannot be implemented in practice because it is not convenient to extract heat from an expanding flow; the turbine construction would be too complicated and not feasible. In addition, due to heat extraction, the moisture at the turbine outlet is far too high (it can reach more than 60%).

What can be done to overcome these construction obstacles is to extract some steam from the turbine at intermediate pressures and condense it to further extract heat for water preheating. This process will diminish turbine work but it will increase the cycle efficiency because the average temperature at source side becomes higher. There are several known ways to transfer heat from the condensing steam extracted from the turbine to the water preheating process. The simplest system is the use of a closed feedwater heater that represents, in fact, a condenser.

In Figure 5.13 is presented a regenerative Rankine cycle with a closed feedwater heater. After it condenses in the regenerator (process 7–8), the working fluid is throttled to lower pressure using a device called a “steam trap” (process 8–9). A fraction f of vapor is extracted from the turbine at intermediate pressure (in state 7) and directed toward the regenerator. The remaining vapor (that is the fraction 1 − f ) is expanded up to condenser pressure (state 10). In a first section of the condenser (10–11) the steam reduces the vapor quality to an intermediate value that corresponds to state 11. The vapor quality of the throttled working fluid in 9 is the same as that in state 11, hence x9 = x11 = x12. The condensation process continues with full flow until saturated liquid is obtained in 1. The relevant mass balance equations and the extracted vapor fraction (f ∈ [0, 1]) for the system are:

m˙6=m˙7+m˙10;fm˙6=m˙7;m˙9+m˙11=m˙12

si284_e  (5.7)

f05-13-9780123838605
Figure 5.13 Regenerative Rankine power plant with closed feedwater heater and steam trap.

The energy balance of the regenerator also imposes that

1-fh7-h8=h3-h2

si285_e  (5.8)

The exergy balance of the regenerator will determine the irreversibility due to heat transfer, owing to the fact that T7 = T8 > T3 > T2. The exergy balance is as follows:

ex2+fex7=ex3+fex8+exd,reg

si286_e  (5.9)

where exd is the exergy destroyed by the regenerator. In addition, there is exergy destruction due to throttling process 8–9. The exergy balance for the steam trap can be written as follows to determine the exergy destruction:

fex8=fex9+exd,trap

si287_e  (5.10)

The energy balance on the condenser requires

fh9+1-fh10=h1+qc

si288_e  (5.11)

where qc represents the mass-specific heat flux rejected by the condenser (in kJ/kg). If m˙si289_e is the mass flow rate of working fluid in state 1, then, with reference to Figure 5.13, it results that the heat rate rejected by the condenser is

Q˙c=m˙qc=Q˙c1+Q˙c2=m˙(qc1+qc2)

si290_e

where qc1 = (1 − f)(h10 − h11) and qc2 = h12 − h1, and one also notes that h9 = h11 = h12.

The vapor quality of the extracted steam in regenerative Rankine schemes is slightly below unity. The efficiency of the regenerative cycle is slightly influenced by the quality of extracted steam. For higher quality, the energy efficiency tends to be better, whereas the exergy efficiency is not influenced. This evolution can be observed in Figure 5.14, which indicates that the best extraction is obtained when vapor quality is around 0.9. It also results that it is detrimental to extract superheated steam because both the energy and exergy efficiencies decrease sharply.

f05-14-9780123838605
Figure 5.14 Energy and exergy efficiency of regenerative steam Rankine cycle as a function of vapor quality of extracted steam [assumptions: cycle configuration as in Figure 5.13; Tc = 30 °C, Tb = 200 °C, T6 = 400 °C, T7 − T3 = 10 °C].

The exergy destructions within the cycle can be reduced and the energy efficiency can be improved if the steam trap is replaced with a pump followed by a mixer. Such a scheme is denoted as a regenerative Rankine cycle with a closed feedwater heater and pump and is represented in Figure 5.15. The system has the advantage of reducing the size of the condenser, and it also reduces the pumping power. However, there are still exergy destructions related to the regenerator and the mixer (the mixer is in fact a direct contact heat exchanger, which, in this case, mixes streams that have about the same temperature). The energy balance on the regenerator becomes in this case:

1-fh2+fh8=1-fh3+fh9

si291_e  (5.12)

whereas the exergy balance becomes

1-fex2+fex8=1-fex3+fex9+exd,reg

si292_e  (5.13)

The energy and exergy balances for the mixer are written as

1-fh3+fh10=h4and1-fex3+fex10=ex4+exd,mix

si293_e  (5.14)

For the ideal cycle exemplified in Figure 5.16 the exergy destruction at the regenerator is 22.86 kJ/kg while the exergy destruction at the mixer is negligible (0.046 kJ/kg). The extracted steam fraction is 11.3% and the energy efficiency of the cycle is slightly improved with respect to the energy efficiency of the one with the steam trap analyzed previously: 35.4% versus 35.1%. This cycle is also illustrated in the P–h diagram in Figure 5.17 where the non-isobaric processes are made apparent.

f05-15-9780123838605
Figure 5.15 Regenerative Rankine power plant with closed feedwater heater, high-pressure pump, and flow mixer.
f05-16-9780123838605
Figure 5.16 Ideal regenerative Rankine cycle with closed feedwater heater of configuration in Figure 5.15 for Tc = 30 °C, Tb = 200 °C, ΔTsh = 200 °C, ΔTreg = 10 °C.
f05-17-9780123838605
Figure 5.17 Ideal regenerative Rankine cycle with closed feedwater heater of configuration in Figure 5.15 for Tc = 30 °C, Tb = 200 °C, ΔTsh = 200 °C, ΔTreg = 10 °C represented in a Th diagram (non-isobaric processes are made apparent).

The exergy destruction in the regenerator–mixer system can be further reduced if the closed feedwater heater is replaced with an open feedwater heater (OFW). This is in fact a direct contact heat exchanger. An improved Rankine plant design of regenerative type is introduced in Figure 5.18. The exergy destruction in the OFW must be smaller with respect to the closed feedwater heater because the temperature of extracted steam can be equal to the temperature of the condensate. In addition, the OFW has a simpler and cheaper construction than the closed one. An example of an ideal cycle with this configuration is illustrated in Figure 5.19.

f05-18-9780123838605
Figure 5.18 Regenerative Rankine power plant with open feedwater heater.
f05-19-9780123838605
Figure 5.19 Ideal steam Rankine cycle with open feedwater heater for Tc = 30 °C, Tb = 200 °C, ΔTsh = 200 °C.

A combination of open and closed feedwater heaters with multiple steam extraction pressures can be used to increase cycle efficiency even more. This approach is used in practice as a technical and economical trade-off for steam Rankine power stations. The OFW is cheaper and more efficient than the closed feedwater heater, but requires additional pumps, which involve more costs. On the contrary, the closed feedwater heaters are more costly and less efficient, but these heaters do not necessitate additional pumps. Moreover, multiple feedwater heaters of closed type can be cascaded using steam traps and a single pump, although this solution destroys more exergy.

Example 5.4

Let us consider a regenerative steam Rankine cycle of the configuration illustrated in Figure 5.13. Assume that Tc = 30 °C, Tb = 200 °C, ΔTsh = 200 °C and that all internal processes are reversible except the feedwater heater, which operates with a temperature difference for heat transfer defined by ΔTreg = T7 − T3 = 10 °C (see Figure 5.13 for the state point numbers). We want to determine the extracted steam ratio and the energy and exergy efficiencies and to represent the cycle in the Ts diagram.

In order to solve this problem mass, energy, entropy, and exergy balance equations must be written for each system component; hence for the nine components of the system, there will be in total 27 equations. Three equations come from the fact that the pump and the turbines operate isentropically, which means that s2 = s1; s6 = s7 = s10. In addition, the extracted steam fraction is defined by Equation (5.8), and because all internal processes are reversible the mechanical and electrical efficiency are ηm = ηel = 1. In total one has 33 equations that form a closed system that can be solved to determine thermodynamic parameters at all of the state points.

A simple code can be written to solve the algebraic system numerically. Here we give the solution obtained with EES as shown in Table 5.9, where the thermodynamic parameters for each of the state points is given. For the specific exergy calculation a reference state must be assumed. As in previous examples, the reference state for the steam Rankine cycle is taken to be liquid water at the standard T0P0. Once the thermodynamic states are determined, the extracted steam fraction results directly from Equation (5.8), which gives the solution f = 12.7%. The energy efficiency is 35.1%. To calculate the exergy efficiency, one assumes that the heat source is at a constant temperature level equal to the temperature of the thermodynamic state at the turbine inlet, Tso = T6 = Tb + ΔTsh = 673 K. Therefore the Carnot factor is 0.5571 and the exergy efficiency is 63.06%. The cycle diagram is presented at scale in Figure 5.20.

Table 5.9

State Point Parameters Calculated for Example 5.4

StateP (kPa)T (K)v (m3/kg)xh (kJ/kg)s (kJ/kgK)ex (kJ/kg)
14.246303.20.0010040125.70.43650.1745
21554303.20.001004-0.3738127.20.43651.73
31554371.20.001041-0.2271411.81.28333.84
41554473.20.0011560852.42.331162.1
51554473.20.1273127936.431879.7
61554673.20.19581.23832557.2521097
7134381.21.2870.998426857.252527.3
8134381.20.0010504531.39741.2
94.246303.24.4320.13474531.5165.573
104.246303.227.970.850321927.25234.25
114.246303.24.4320.13474531.5165.573
124.246303.24.4320.13474531.5165.573

t0050

f05-20-9780123838605
Figure 5.20 Ideal regenerative Rankine cycle for steam power plant with closed feedwater heater and steam trap for Tc = 30 °C, Tb = 200 °C, ΔTsh = 200 °C, ΔTreg = 10 °C (Example 5.4).

Example 5.5

In Figure 5.21 is exemplified a steam Rankine power plant with multiple feedwater heaters and steam extraction points. The cycle operates according to the following main parameters: Tc = 30 °C, Tb = 200 °C, ΔTsh = 200 °C.

f05-21-9780123838605
Figure 5.21 Regenerative Rankine power plant configuration with multiple feedwater heaters.

Mass balance equations for this system become significantly more complex with respect to the other systems discussed previously. Table 5.10 gives the mass balance equations for all components of the system and also introduces the steam extraction fractions fi, i = 11, 14, 17, 18. Using the steam fractions the energy and exergy balances can be written for each plant component according to Equations (5.8)(5.14).

Table 5.10

Mass Balance Equations for Rankine Cycle Represented in Figure 5.21

ComponentMBEComponentMBEf definitions
LPPm˙2=m˙1si55_eCFWH2–3m˙3=m˙2;m˙18=m˙19=m˙20si56_em˙11=f11m˙10si57_e
HPPm˙5=m˙4si58_eOFWHm˙3+m˙16+m˙17=m˙4si59_em˙14=f14m˙10si60_e
ST15–16m˙16=m˙15si61_eCFWH5–6m˙6=m˙5;m˙13+m˙14=m˙15si62_em˙17=f17m˙10si63_e
PHm˙8=m˙7si64_eCFWH6–7m˙7=m˙6;m˙11=m˙12si65_em˙18=f18m˙10si66_e
Bm˙9=m˙8si67_eSTBf11 + f14 + f17 + + f18 + f21 = 1m˙21=f21m˙10si68_e
SHm˙10=m˙9si69_eCONDm˙20+m˙21=m˙1si70_em˙1=f18+f21m˙10si71_e

t0055

MBE, Mass balance equation; LPP, low-pressure pump; CFWH, closed feedwater heater; OFWH, open feedwater heater; HPP, high-pressure pump; ST, steam trap; PH, preheater; B, boiler; SH, superheater; STB, steam turbine; f, steam extraction fraction.

Optimal flow extraction fractions can be determined based on various multivariable optimization methods to maximize the energy efficiency of the cycle. Here the genetic algorithm optimization implemented in EES is used to maximize cycle efficiency; the optimum extraction fractions are determined as follows: f11 = 5.4%, f14 = 4.7%, f17 = 5.8%, f18 = 3.8%. The maximized energy efficiency for the ideal Rankine cycle case study is η = 36.5%. The exergy efficiency of the cycle is 65.5%; exergy is destroyed in feedwater heaters, steam traps, the condenser, and the steam generator. The thermodynamic cycle is represented in the Ts diagram shown in Figure 5.22.

f05-22-9780123838605
Figure 5.22 Ideal steam Rankine cycle for the power plant represented in Figure 5.21 [assumptions:Tc = 30 °C, Tb = 200 °C, ΔTsh = 200 °C].
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset