5.2.5 Steam Rankine Power Stations

5.2.5.1 Reheating-Regenerative Steam Rankine Cycle

The vast majority of power stations are based on reheating-regenerative steam Rankine cycles. These types of cycles are very complex as they comprise many types of components, including preheater, boiler, superheater, reheater, closed (CFWH) and open (OFWH) feedwater heaters, mixer, condenser, deaerator, pumps, and steam traps (ST). The cycle design depends on the type of heat source and heat sink. For most cases of large power generation the heat source is derived from coal combustion.

The heat sink is in general a lake; however, there are many cases in which power plants use air as a heat sink. In this case the plant is equipped with cooling towers—either with natural convection or with forced convection. In some situations (e.g., in colder regions) the heat sink is represented by a district heating system for cogeneration. In this section we give an illustrative example of a reheating-regenerative steam Rankine cycle as presented in Figure 5.23. This power plant comprises eighteen main components. There are also four steam extraction ports with corresponding steam extraction fractions denoted with f12, f17, f20, f21. The thermodynamic analysis of the cycle becomes complex because multiple variables and equations are involved and optimization is required for energy efficiency maximization. We will illustrate in detail how the cycle can be analyzed using balance equations and how the optimization problem for maximum efficiency can be formulated.

f05-23-9780123838605
Figure 5.23 Rankine power plant with single reheating and multiple feedwater heaters (M, mixer; CFWH, closed feedwater heater; OWFH, open feedwater heater; ST, steam trap).

Mass, energy, entropy and exergy balance equations must be written for each component of the power plant in order to form a system of equations that can be solved to determine thermodynamic parameters defining each state. The following observation is very useful for simplifying the mass balance equations: in state 11 (turbine inlet) the mass flow rate is the maximum for the plant; also m˙11=m˙10=m˙9=m˙8si294_e. Hence, the steam injection fractions are defined with respect to m˙11si295_e as follows:

f12m˙11=m˙12,f17m˙11=m˙17,f20m˙11=m˙20,f21m˙11=m˙21

si296_e

In addition, in Figure 5.23 it can be observed that mass balance equations can be reduced to the following set of equations expressed as a function of flow fractions in other state points of the power plant, defined by fim˙11=m˙isi297_e, namely:

f1=f2=f3=f26;f4=f5=f6=f7;f12=f13=f14;f17=f18=f19;f21=f22=f23

si298_e

f15=f16;f24=f25;f25+f23=f26;f12+f15=1;f4=f15;f16=f17+f20+f21+f24

si299_e

Using the flow fractions fi, the energy, entropy, and exergy balance equations are written for each component as indicated in Table 5.11. The balance equations together with the parameters that quantify exergy destructions (e.g., isentropic efficiencies of pumps and turbines, temperature differences due to heat transfer) form a closed system of equations which can be solved for temperatures, pressures, and specific enthalpy, entropy, and volume at each state point provided that all extraction fractions are specified (f12f17f20f21) and the condensation, boiling, and superheating temperatures (T1T9T11) are given.

Table 5.11

Balance Equations for Ideal Rankine Cycle of Power Plant in Figure 5.23

ComponentEquationsComponentEquations
Pump 1 (states 1–2)MBE: m˙1=m˙2si72_e
EBE: f1h1 + wp1 = f1h2
EnBE: s1 + sg = s2
ExBE: f1ex1 + wp1 = f1ex2 + exd
CFWH1 (states 2–3)MBE: m˙2=m˙3si73_e
EBE: f1h2 + f21h21 = f1h3 + f21h22
EnBE: f1s2 + f21s21 + sg = f1s3 + f21s22
ExBE: f1ex2 + f21ex21 = f1ex3 + f21ex22 + exd
Pump 2 (states 4–5)MBE: m˙4=m˙5si74_e
EBE: f4h4 + wp2 = f4h5
EnBE: s4 + sg = s5
ExBE: f4ex4 + wp2 = f4ex5 + exd
OFWH (states 3–4–19–20)MBE: f1m˙11+f17m˙11+f20m˙11=f4m˙11si75_e
EBE:f1h3 + f17h18 + f20h20 = f4h4
EnBE: f1s3 + f17s18 + f20s20 + sg = f4s4
ExBe: f1ex3 + f17ex18 + f20ex20 = f4ex4 + exd
Pump 3 (states 13–14)MBE: m˙13=m˙14si76_e
EBE: f12h13 + wp3 = f12h5
EnBE: s13 + sg = s14
ExBE: f12ex13 + wp3 = f12ex13 + exd
CFWH2 (states 5–6–17–18)MBE: f4m˙11+f17m˙11=f4m˙11+f17m˙11si77_e
EBE: f4h5 + f17h17 = f4h6 + f17h18
EnBE: f4s5 + f17s17 + sg = f4s6 + f17s18
ExBE: f4ex5 + f17ex17 = f4ex6 + f17ex18 + exd
Mixer (states 7–8–14)MBE: f4m˙11+f12m˙11=f8m˙11si78_e
EBE: f4h7 + f12h14 = h8
EnBE: f4s7 + f12s14 + sg = s8
ExBE: f4ex7 + f12ex14 = ex8 + exd
CFWH3 (states 6–7–12–13)MBE: f4m˙11+f12m˙11=f4m˙11+f12m˙11si79_e
EBE: f4h6 + f12h12 = f4h7 + f12h13
EnBE: f4s6 + f12s12 + sg = f4s7 + f12s13
ExBE: f4ex6 + f12ex12 = f4ex7 + f12ex13 + exd
Preheater (states 8–9)MBE: m˙8=m˙9si80_e
EBE: h8 + qph = h9
EnBE: s8 + qph/Tph + sg = s9
ExBE: ex8 + qph(1 − T0/Tph) = ex9 + exd
Boiler (states 9–10)MBE: m˙9=m˙10si81_e
EBE: h9 + qb = h10
EnBE: s9 + qb/Tb + sg = s10
ExBE: ex9 + qb(1 − T0/Tb) = ex10 + exd
Superheater (states 10–11)MBE: m˙10=m˙11si82_e
EBE: h10 + qsh = h11
EnBE: s10 + qsh/Tsh + sg = s11
ExBE: ex10 + qsh(1 − T0/Tsh) = ex11 + exd
Reheater (states 15–16)MBE: m˙15=m˙16si83_e
EBE: f15h15 + qrh = f15h16
EnBE: f15s15 + qrh/Tsh + sg = f15s16
ExBE: f15ex15 + qrh(1 − T0/Trh) = f15ex16 + exd
Turbine 1 (states 11–12–15)MBE: m˙11=m˙12+m˙15si84_e
EBE: h11 = f12h12 + f15h15 + wt1
EnBE:s11 + sg = f12s12 + f15s15
ExBE: ex11 = f12ex12 + f15ex15 + wt1 + exd
Turbine 2 (states 16–17–20–21–24)MBE: f15 = f17 + f20 + f21 + f24;
EBE: f15h16 = f17h17 + f20h20 + f21h21 + f24h24 + wt2;
EnBE: f15s15 + sg = f17s17 + f20s20 + f21s21 + f24s24;
ExBE: f15ex15=f17ex17+f20ex20+f21ex21+f24ex24+wt2+exdsi85_e
ST 1 (states 22–23)MBE: m˙22=m˙23si86_e
EBE: h22 = h23
EnBE: f21s22 + sg = f21s23
ExBE: f21ex22 = f21ex23 + exd
ST 2 (states 18–19)MBE: m˙18=m˙19si87_e
EBE: h18 = h19
EnBE: f17s18 + sg = f17s19
ExBE: f17ex18 = f17ex19 + exd

t0060

As seen in Table 5.11 the balance equations can be written using mass-specific quantities. These quantities are defined with respect to the mass flow rate at the HPT inlet; thus: w=W˙/m˙11si300_e and q=Q˙/m˙11si301_e. Furthermore, the energy efficiency can be maximized according to the following objective function:

maxη=wnetqin=Ff12f17f20f21withgivenT1,T9,T11andfi01

si302_e

where F(f12f17f20f21) represents the functional relation between steam extraction fractions and the energy efficiency of the cycle, a relation which is established by the nonlinear system of equations mentioned above.

Note that the certain additional assumptions must be made in order for the optimization problem to be completely specified. These assumptions are as follows:

 at pump suction the liquid is saturated, hence x1 = x4 = x13 = 0

 at the steam traps inlet the liquid is saturated, hence x18 = x22 = 0

 at the OFW outlet there is a saturated liquid, hence x4 = 0

 there is a minimum temperature difference at the closed feedwater heater outlet denoted with ΔThx; it is assumed that T22 − T3 = T18 − T6 = T13 − T7 = ΔThx.

Example 5.6

Let us consider a steam cycle of the configuration described in Figure 5.23. The following parameters are assumed: Tc = 30 °C; Tb = 293 °C; Tsh = Trh = 400 °C; ΔThx = 10°C, where Tsh and Trh are the superheating and reheating temperatures, respectively. We want to determine the optimum steam extraction fractions that maximize cycle energy efficiency for two cases: (i) all pumps and turbines operating reversibly and (ii) the isentropic efficiency of the pumps and steam turbines is 0.85. Exergy destructions for each of the components will be determined for case (ii).

The results of the optimization problem (solved with EES software) are presented in Table 5.12 for case (i). We assume that water is saturated liquid at the pump inlet and at closed feedwater heater outlets (states 13, 18, 22). There are optimal values for steam extraction fractions and optimum extraction pressures. It can be observed that the largest extraction is required after the first stage of expansion with f12 = 14.4% and the extraction fraction decreases when decreasing the pressure. The mass-specific net generated work is 1069 kJ/kg while heat input is 2350 kJ/kg.

Table 5.12

Performance Parameters of Optimized Ideal Rankine Cycle of Figure 5.23

ParameterValue
Parameters with assumed valuesTc = 30 °C; Tb = 293 °C, Tsh = Trh = 400 °C; ΔThx = 10 °C
Optimum extraction fractionsf12 = 14.4%f17 = 6.7%f20 = 8.3%f21 = 5.1%
Optimized extraction pressuresP12 = 3, 381 kPa; P17 = 988 kPa; P20 = 266 kPa; P21 = 39.3 kPa
Energy and exergy efficiencies and Carnot based efficiencyη=wnetqin=45.5%,ψ=η1T0/Tsh=81.6%si88_e
Net generated workwnet = wt1 + wt2 − wp1 − wp2 − wp3 = 1, 069 kJ/kg
Heat input and exergy consumedqin = h11 − h8 = 2, 350kJ/kg; excons = ex11 − ex8 = 885kJ/kg
Exergy destroyedexd = ∑ exd = 62.18 kJ/kg; for each component the exergy destruction in percents with respect to total is as follows: exd,c = 34%; exd,CFWH1 = 12.6%
exd,CFWH2 = 11.2%; exd,CFWH3 = 19.9%; exd,M ≅ 0.2%;
exd,OFWH = 20.1%; exd,ST1 = 1.0%; exd,ST2 = 1.0%
Pressure ratio over turbinesPR1 = P11/P12 = 2.3; PR2 = P16/P24 = 796
Expansion ratio over turbineER1=v12v11=1.91;ER2=v24v16=300si89_e
Back work ratioBWR=wp1+wp2+wp3wt1+wt2=0.7%si90_e

Note: Tc = T1Tb = T9Tsh = Trh = T11.

The energy efficiency (which is maximized) reaches 45.5%. Moreover, one notes that vapor quality at the isentropic turbine outlet is 80%, which is acceptable. The exergy efficiency of the system is exceptionally good, namely ψ = 81.6%, due to the assumed reversible operation of pumps and turbines. Note that the average temperature at the steam generator is approximated to 603.2 K (Carnot factor 50.6%); the Carnot factor associated with the turbine inlet temperature is 55.7%.

The total exergy destruction for case (i) is 61.18 kJ/kg. The highest exergy destruction occurs in the condenser at 34% of total, followed by the OFW and the high-pressure closed feedwater heater CFWH3, each with ~ 20% of total exergy destruction. The exergy destruction in the mixer is negligible.

To solve the cycle in case (ii) the assumption that water is saturated liquid at the pump inlet and at closed feedwater heater outlets is maintained. In addition, the assumption that there is negligible pressure drop in all devices with heat exchange and in pipelines is maintained. Furthermore, for case (ii) temperature differences are assumed for the steam generator as follows: 75 °C for the preheating section, 50 °C for the boiling section, and 25 °C for the superheating and reheating sections.

In these conditions the optimum extraction rates are found to have relatively similar values as for case (i), as follows: f12 = 13.3%, f17 = 6.1%, f20 = 7.7%, f21 = 4.8%. The pressure ratio over the HPT is 2.8 and over the LPT is 660. All state point parameters of the optimized cycle are given in Table 5.13. The energy efficiency for case (ii) is degraded to 39.2%, as compared to the case (i) with 45.5%.

Table 5.13

Cycle Parameters of Reheating-Regenerative Steam Rankine Cycle for Case (ii)

StateP (kPa)T (K)v (m3/kg)x (kg/kg)h (kJ/kg)s (kJ/kgK)ex (kJ/kg)Mass-specific parameters (kJ/kg)
Componentwqexd
14.246303.20.0010040125.70.43650.1745Pump 10.3N/A0.04
2244.6303.20.001004− 0.18621260.43670.4166CFWH1N/A1088.05
3244.6338.60.00102− 0.1183274.10.898810.77OFWHN/A26212.0
4244.6399.90.0010670532.41.660.07Pump 28.2N/A0.91
57769400.90.001063− 0.5245541.81.60368.44CFWH2N/A1366.83
67769437.41.977− 0.417698.41.977113.6CFWH3N/A25611.2
77769503.60.001203− 0.2144993.32.604221.5MN/AN/A~ 0.0
87769503.80.001203− 0.2138994.12.606221.8Preheat.N/A31221.3
97769566.20.001377013053.187359.7BoilerN/A145662.2
107769566.20.02433127615.7591049Superheat.N/A3817.3
117769673.20.035491.26231436.3821245HPT213N/A20.8
122820546.70.0811910029306.4511011ReheaterN/A2645.1
132820503.60.001210992.32.614217.5LPT743N/A125.3
147769504.90.001205− 0.2102999.32.616223.9CondenserN/A146624.2
152820546.70.081191.0729306.4511011ST1N/A0.00.62
162820673.20.1061.23832346.9531165ST2N/A0.00.54
17877535.30.27361.09829737.041878.4Cycle parameters
18877447.40.001120738.12.084121.4Heat inputqin = 2, 413 kJ/kg
19244.6399.90.07010.09422738.12.114112.4Net workwnet = 946 kJ/kg
20244.6419.30.77371.01927577.161626.7Efficiencyη = 39.2%
2139.28348.63.8450.946525117.319333.8Exergy inputexin = 1, 253 kJ/kg
2239.28348.60.0010260315.81.02116.1Efficiencyψ = 75.5%
234.246303.22.5750.07825315.81.0643.31Exergy destructionexd = 306.6 kJ/kg
244.246303.229.050.88322717.51335.56Pressure ratioHPT: 2.8; LPT: 660

t0070

This example underlines the importance of irreversibilities in the steam generator and the turbines. Due to the additional exergy destruction the cycle exergy efficiency for case ii) reduces to 75.5% (for case (i)) the exergy efficiency is 81.6%). The relative exergy destructions for cycle components are represented in the chart in Figure 5.24. Under the assumed conditions the most exergy destruction occurs in the LPT (41%), followed by the steam generator (20%).

f05-24-9780123838605
Figure 5.24 Exergy destructions for reheating-regenerative cycle in Example 5.6 (case ii).

5.2.5.2 Coal-Fired Power Stations

Although their flue gas stacks highly pollute the atmosphere with carbon dioxide, sulfur dioxide, nitrous oxide emissions, and other pollutants, coal-fired power stations are used in nearly every country of the world. These types of power generating stations are responsible for the majority of global electricity production. They operate based on steam Rankine cycles (generally on reheating-regenerative schemes) and comprise a coal-fired furnace as steam generator.

Two main combustion technologies are used currently in coal-fired power stations, namely pulverized coal boilers (the most predominant) and fluidized bed boilers (used when the coal rank is lower). Although the two technologies are different with regard to transport phenomena, the transient heat and mass transfer from a combusting coal particle to the surroundings are similar in both cases: combustion occurs at the coal particle surface, which gradually consumes while generates incandescent, hot gases. Coal is supplied at the lower part of the furnace where it devolatilizes and ignites in an oxidative atmosphere. The oxygen comes from two sources: fresh air injection and recirculated flue gases. In the first part of the combustion process, the volatiles are oxidized while the temperature of coal particle increases. Thereafter, fixed carbon oxidizes at very fast rate.

In Figure 5.25 the typical configuration of a pulverized coal-fired furnace at a conventional power station is illustrated. It has the roles of preheating, boiling, superheating, and reheating high-pressure steam. The furnace normally is in the form of a tall brick structure of a 10m to 50m height, depending on the type of coal and the residence time of coal particles required for complete combustion. Pulverized coal-fired furnaces are crucial to the utility industry. Typically, pulverized coal-fired furnaces comprise a coal mill which grinds coal to form small powder-like particles conveyed with primary air to the pulverization point, where ignition occurs.

f05-25-9780123838605
Figure 5.25 Typical coal-fired steam generator configuration for conventional power stations.

The average temperature of gases at the lower part of the pulverized coal furnace (below point A in Figure 5.25) is typically around 650–700 K. A calandria boiler system with vertical tubes is installed at the lower half of the furnace. Between the hot gases inside the furnace duct and the calandria tubes placed at the periphery an intense radiative heat transfer occurs. Under exposure to intense thermal radiation, water boils.

The calandria system comprises a steam drum (see Figure 5.25) which accumulates liquid water at the lower part. Water flows gravitationally downward and feeds into a system formed of multiple headers arranged such that water is distributed uniformly into calandria tubes. The boiling process in the calandira system is incomplete and is facilitated by the fact that the water is close to its saturation state. Given that the boiling water is lighter than the feed water, in the calandria system is formed a natural circulation process due to gravity. In some cases circulation pumps are also used to enhance the water circulation rate.

In later stages of the combustion process, carbon oxidizes and the flue gas temperature increases further. This process occurs approximately between locations A and B in Figure 5.25. At point B, coal particles are completely consumed and the flue gas temperature reaches its maximum value (around 1200 K). In between points B and C the heat of the flue gases is transferred (mainly radiatively) to the superheater and reheaters of the steam Rankine plant. At points C and D there is a heat transfer process by forced convection and radiation between flue gases and the boiling water prior its return to the steam drum. Saturated vapor exists above the liquid in the steam drum. The next heat transfer process in the furnace occurs between D and E where the economizer is installed, which preheats water prior to feeding it into the steam drum. The heat transfer in the economizer is mostly caused by a forced convection process. After point E, a part of the oxygen-containing hot flue gases is returned to the combustion zone in the furnace. Fly ash is removed and electrostatic devices are typically used to remove particulate matter. Further, the flue gases are directed to the stack using a mechanical draft system.

An example of Rankine cycle configuration for a coal-fired steam Rankine power station is shown in Figure 5.26. The system is a double reheating-regenerative Rankine cycle. Liquid water in 1 is pumped by a low-pressure pump (LPP), preheated in a low-pressure closed water-feeding heater (LPFW), further heated in the OFW, and then is pressurized by the high-pressure pump (HPP).

f05-26-9780123838605
Figure 5.26 Diagram of a coal-fired steam Rankine cycle.

After state 5, pressurized water is the preheated in closed feedwater heater (HPPFW) and in the economizer, and then it is fed as saturated liquid in 7 to the steam drum. Saturated vapor is extracted from the steam drum in 8 and superheated and expanded in the HPT. The expanded steam in 10 is then reheated (RH1) and expanded in the intermediate-pressure turbine (IPT). The expanded steam in 12 is again reheated (RH2) and expanded in the LPT. Steam is extracted in 15, 17, 18, and 19 to be used in the water preheating process. In the calandria system, saturated liquid water from the steam drum bottom is extracted in 22 and heated radiatively in the “Calandria HXs” and then further heated by combined convection–radiation heat transfer in the “BOILER” section. The figure illustrates the heat transfer between hot flue gases and steam as shown with bold line. The process B–C represents an approximation of the heat transfer process between the flue gases and the steam at the superheater and the two reheaters. The process C–D represents the heat transfer in the “BOILER” heat exchanger (see Figure 5.26), whereas process D–E represents the heat transfer in the preheater.

The coal type and its specific energy and exergy contents play a role mostly in the design of the furnace and the efficiency of the power plant as well as its air pollution characteristics. The gross calorific value of coal varies from 13 to ~34 MJ/kg; chemical exergy (which is sensitive to moisture content) varies from about 5 to ~25 MJ/kg; and the specific GHG emissions of coal combustion are in the range of 70–140 kg CO2 equivalent per GJ on a net calorific value basis (see Chapter 3).

For thermodynamic analysis the mass, energy, entropy, and exergy balance must be written for each component of the system. A logical sequence of computation steps requires solving the equations related to coal combustion in the furnace first.

In the actual coal furnace system oxidant in excess is used. Furthermore, additional oxygen is supplied to the combustion process from recirculated flue gases. Therefore, the combustion process can be represented by the following equation

R+yP+λxstO2+3.76N21+yP+aA

si303_e  (5.15)

where Rsi304_e stands for reactant (referring to coal only) and Psi305_e for products (except the ash), respectively, xstsi306_e represents the stoichiometric molar fraction of fresh oxygen, λ > 1 is the excess air factor, y0si307_e is the flue gas recycling factor, and Asi308_e stands for the ash. It is useful to describe coal using atomic fractions expressed per mole of coal on an as-received basis. Hence, the reactants Rsi309_e can be represented by the following equation

R=cC+0.5hH2+sS+0.5oO2+0.5nN2+aA+mH2O

si310_e  (5.16)

where c,h,s,o,n,a,msi311_e are determined from weight fractions as-received by dividing with molecular masses.

The reaction products, assuming that complete consumption of coal generates only the products CO2, H2O, SO2, O2, N2, and ash, is described by the following equation:

P=cCO2+0.5h+mH2O+sSO2+λ-1xstO2+0.5n+3.76λxstN2

si312_e  (5.17)

The stoichiometric molar fraction of fresh air, xstsi313_e, results from atomic species balance:

xst=c+0.25h+s-0.5o

si314_e  (5.18)

As mentioned above, the first phase of the combustion process is represented by the process of pyrolysis coal and volatile combustion and occurs in a spatial region that extends approximately between the furnace bottom and the coal’s pulverization location. Figure 5.27 gives a sketch representing the matter and heat fluxes in the ignition and volatile combustion zone of the furnace. During the pyrolysis, a series of combustible materials such as long- and short-chain hydrocarbons (e.g., CnH2n + 2) and sulfur and some amount of carbon dioxide are released. Hence, a part of fixed carbon is released during pyrolysis as hydrocarbons with a carbon-to-hydrogen ratio of ~0.25–0.47. If one denotes with z01si315_e the fraction of fixed carbon released during pyrolysis, then the volatile combustion phase can be described by the following chemical equation

R+λxstO2+3.76N2+yPc1-zC+Pvc+aA

si316_e  (5.19)

where Pvcsi317_e represents the combusted volatile gases defined according to

Pvc=nH2H2+nCO2CO2+nH2OH2O+nSO2SO2+nO2O2+nN2N2

si318_e  (5.20)

with

nH2=0.5h1-z;nCO2=cz+y;nH2O=0.5hz+y+m1+y;nSO2=s1+y

si319_e

nO2=1-zc+0.25h+1+yλ-1xst;nN2=1+y0.5n+3.76λxst

si320_e

standing for the number of moles of components per mole of combusted coal.

f05-27-9780123838605
Figure 5.27 Representation of ignition zone process and involved streams (furnace bottom).

In the second phase of combustion (see Figure 5.28) the remaining un-combusted fixed carbon from the first phase oxidizes and the temperature of flue gases increases. Recirculated gases are added to the process as a means of adjusting the temperature at the end of combustion phase. The combustion process (see A–B in Figure 5.25) is described by the following equation

Pvc+c1-zC1+yP

si321_e  (5.21)

f05-28-9780123838605
Figure 5.28 Representation of fixed carbon combustion process and involved streams.

Notice that the summation of Equation (5.19) and Equation (5.21) gives Equation (5.15), which describes the overall process of coal combustion.

The boiling process continues in the heat exchanger placed between points C and D as indicated in Figure 5.25. Once the thermodynamic states in the furnace (0, A, B, C, D, E, F, 6, 7, 8, 9, 10, 11, 22, 23, 24, 25) are completely determined the analysis proceeds further with balance equations of all other components of the Rankine steam power plant. According to energy balance, the mass flow rate in states 4, 5, 6, 7, 8, 9, and 10 must be the same. According to the overall energy balance on the boiler one has

Q˙boiler=H˙25-H˙22=m˙7h8-h7

si322_e  (5.22)

where H˙=m˙hsi323_e is the enthalpy rate of the flow.

In the above equation one accounts for the fact that a saturated liquid enters in state 7 whereas a saturated vapor exits out of the drum in 8. The equation is then solved for m˙7si324_e. The enthalpy in 25 results from the enthalpy in state 24 and the enthalpies of states C and D on the flue gas side, according to the energy balance

H˙25-H˙24=H˙C-H˙D

si325_e

In the example presented above the NOx emissions are neglected, which is a reasonable assumption for first hand calculations. More detailed models must include chemical equilibrium analysis of the combustion process and chemical kinetics in order to determine the NOx emissions. As a rule of thumb, NOx emissions represent half of SO2 emissions per unit of kW net power. On average, for a coal-fired power plant the specific emissions are as follows: 900 g/kWh CO2, 7 g/kWh SO2, and 4 g/kWh NOx.

In the subsequent part of this section the summary of two case studies of energy and exergy analyses of two actual power plants is given. A larger-scale power plant is selected, as Nanticoke Power Generation Station (NPGS) in Ontario, with 500-MW installed capacity per power generating unit. The exergy analysis of the NPGS power plant is summarized here based on the work of Dincer and Rosen (2013). For comparison purposes, the second case study refers to a lower-capacity power plant with a 32-MWgenerator output, operated by Tecro Power Systems Ltd. (TPSL) in India which was previously analyzed in Regualagadda et al. (2010).

The NPGS has steam-reheating power generating units with three stages of turbines (high, intermediate, and low pressure) and a water-cooled condenser. The steam generator operates on the coal pulverization principle. The feedwater system comprises LPP and HPP, two closed feedwater heaters with steam traps, one OFW, and five steam extraction points from the turbines.

The TPSL power plant has a circulating fluidized bed furnace as a steam generator with a nominal capacity of 140 t/h superheated steam. The ultimate coal analysis for TPSL is as follows: moisture 25%, ash 0.88%, hydrogen 4.06%, nitrogen 1.1%, sulfur 0.075%, oxygen 7.935%, carbon 60.95%. The feedwater system comprises a LPP, a HPP, three closed feedwater heaters, and one OFW. The steam cycle operates with no reheating and the turbine has three steam extraction points. The condenser is aircooled.

The general technical parameters of both power stations are compared in Table 5.14. The layout of a Rankine plant is described by the simplified diagram in Figure 5.29, which is applicable to both the NPGS and TPSL plants and consists of four main subsystems: feedwater system, steam generator, power production system, and condenser. The state points indicated in the diagram are described in Table 5.15, and the energy and exergy are given for each stream. In addition, the exergy destructions are given in the Table 5.16. As remarked in Table 5.14, the coal quality is much better at NPGS, hence the efficiencies are higher.

Table 5.14

General Technical Parameters of NPGS and TPSL Power Stations

ParameterNPGSTPSL
Cycle typeSingle reheatingNo reheating
Nameplate power523 MW32 MW
Coal GCV32.8 MJ/kg22.0 MJ/kg
Coal exergy34.2 MJ/kg26.0 MJ/kg
Steam generatorPulverized coalCirculating fluidized bed
Condenser typeWater cooledAir cooled
Sink temperature287 K314 K
Energy efficiency37%25%
Exergy efficiency36%30%
Specific emissions1.262 kg GHG/kWh1.218 kg GHG/kWh
f05-29-9780123838605
Figure 5.29 Simplified layout of steam Rankine power generating station (state point descriptions in Table 5.14).

Table 5.15

Streams Parameters for NPGS (523 MW) and TPSL (32 MW) Generating Units

StreamDescriptionNPGSTPSL
T (K)E˙si91_e (MW)Ex˙si92_e (MW)T (K)E˙si93_e (MW)Ex˙si94_e (MW)
1Coal fed30013681427300120124
2Air3000030000
3Flue gas3937462500269
4High-pressure steam710178083879312546
5Heat loses from high-temperature steam29810029840
6Net powern/a524524n/a3232
7Saturated vapor30877554331671.6
8Extracted steam55044914249775.6
9Condensed vapors30826033170
10Heat rejected by the condenser3087460331600
11Pump power inputn/a1313n/a11
12Preheated water500487132473314

t0080

Table 5.16

Exergy Destruction of Power Plant Subsystems for NPGS and TPSL

ComponentNPGSTPSL
Feedwater system23 MW (3%)2.6 MW (3%)
Steam generator600 MW (75%)73 MW (87%)
Power production118 MW (15%)6.9 MW (8%)
Condenser54 MW (7%)1.6 MW (2%)
Total795 MW (100%)84.1 MW (100%)

Due to a lower quality of coal the exergy destructions of the TPSL steam generator are relatively more important than at NPGS (87% vs. 75% with respect to overall exergy destructions). Comparatively, the feedwater system destroys 3% of exergy for both the NPGS and TPSL stations. The destruction of exergy in the power production subsystem is more important at NPGS than at TPSL (15% vs. 8%).

Conventional power generators based on a the steam Rankine cycle using fossil fuels other than coal (e.g., natural gas or fuel oil) do exist. However, due to economic advantages and efficiency improvements, oil- and gas-fired power plants are combined types, including a combustion turbine and a steam turbine. These types of combined plants will be discussed later in this chapter, after the introduction of gas turbine power plants below.

Example 5.7

We give an example of thermodynamic analysis of a coal-fired power plant having a furnace arranged according to the sketch in Figure 5.25 and a steam Rankine cycle according to Figure 5.26. Assume that the coal power plant operates with good quality coal, anthracite, with ultimate analysis as indicated in Table 5.17.

Table 5.17

Coal and Ash Composition from Ultimate Analysis, By Weight, As Received

Ultimate analysis (% by weight AR)Ash analysis (% by weight)
Carsi95_earSarsi96_eOarsi97_eNarsi98_eAarsi99_eMarsi100_eMcoalSO3K2OTiO2MgOCaOFe2O3Al2O3SiO2Mash
757.515128.56.8243211030252585.02

t0090

Mcoal, molecular mass of coal in kg/kmol; Mash, molecular mass of ash in kg/kmol; molecular mass of compound (coal or ash) is calculated based on component weight fractions wi and molecular weight of components Mi with the equation: M = (∑ wi/Mi)− 1; AR, as-received.

The chemical exergy, formation enthalpy, and formation entropy for coal are calculated first, based on equations indicated in Chapter 3. In Table 5.18 the calculation sequence for coal composition (such as dry-basis, dry, ash-free basis, and atomic fractions of constituents) and the main thermodynamic parameters of coal and ash (such as chemical exergy, calorific values, formation entropy, and enthalpy) are given. Based on the assumed coal and ash composition the value of 99 or 13.2 MJ/kgar is obtained for chemical exergy on an as-received basis.

Table 5.18

Coal Compositions and Estimated Chemical Exergy, Formation Enthalpy and Entropy

ParametersEquationValue
Dry basis ultimate analysis (kg/kgdb)CdbCar=dbar=SdbSar=OdbOar=NdbNar=AdbAar=100100Marsi101_e82.0, 8.2, 1.1, 5.5, 1.1, 2.2a
Dry ash-free ultimate analysis (kg/kgdaf)CdafCdb=dafdb=SdafSdb=OdafOdb=NdafNdb=100100Adbsi102_e83.8, 8.4, 1.1, 5.6, 1.1b
Atomic fractions in daf basis (kmol/kgdaf)cdaf=Cdaf1200;hdaf=Hdaf100;sdaf=Sdaf3200;odaf=Odaf1600;ndaf=Ndaf1400si103_e69.8, 83.8, 0.3, 3.5, 0.8c
Atomic fractions in as-received basis (kmol/kgar)car=Car1200;har=Har100;sar=Sar3200;oar=Oar1600;nar=Nar1400;si104_e

aar=Aar101,200;mar=Mar1800si105_e
62.5, 75, 0.31, 3.12, 0.71, 0.23, 4.7d
Atomic fractions per kmol of coal as received (kmol/kmolar)c=carMcoal;h=harMcoal;s=sarMcoal;o=oarMcoal;n=narMcoalsi106_e
a=aarMcoal;m=marMcoalsi107_e
0.426, 0.512, 0.002, 0.021, 0.005, 0.002, 0.032e
Formation entropy of ashΔSash0 = ∑ wiΔSi0 in kJ/kg.K, Table 3.5 for formation entropies of components. Sash0 = ΔSash0Mash in kJ/kmol.K0.66 kJ/kg.K
56 kJ/kmol.K
Formation enthalpy of ashΔHash0 = ∑ wiΔHi0, in MJ/kg, Table 3.5 for formation enthalpies of components. Hash0 = ΔHash0Mash in kJ/kmol− 11.4 MJ/kg
− 968 MJ/kmol
Chemical exergy of ashexashch = ∑ wiexich in MJ/kg, Eq. (1.41), Table 1.6, Table 3.4 for ΔG0 = ΔH0 + T0ΔS0exashch,mol = exashchMash in MJ/kmol0.4878 MJ/kg
41.5 MJ/kmol
Coal formation entropyΔSdaf0 is calculated according to Equation (3.11) in dry, ash-free basis55.1 kJ/kg.K
Coal formation entropy in as-received basisΔSar0kJ/kg.K=100AarMar100ΔSdaf0+Aar100ΔSash0+Mar100ΔSH2O0;si108_e

Sar0[kJ/kmol. K] = ΔSar0Mcoal
49.66 kJ/kg.K
338.7 kJ/kmol.K
Gross calorific valueGCVdb is calculated based on Equation (3.8); GCVdaf=100/100AdbGCVdbsi109_e37.9 MJ/kgdb
38.8 MJ/kgdaf
Coal formation enthalpyΔHdaf0=GCVdaf+cΔHCO20+hΔHH2O0+sΔHSO20;ΔHar0kJ/kgar=100AarMar100ΔHdaf0+Aar100ΔHash0+Mar100ΔHH2O0;Har0kJ/kmol=ΔHar0Mcoalsi110_e− 0.75 MJ/kgdaf
− 2.25 MJ/kgar
− 15 MJ/kmol
Coal chemical exergyexdafch is calculated with Equation (3.12);

exarch=100AarMar100exdafch+Aar100exashch+Mar100exH2Och;si112_e

exarch,mol = exarchMcoal
22.3 MJ/kgdaf
19.9 MJ/kgar
136.1 MJ/kmol

db, dry basis; ar, as received; daf, dry, ash free.

a Dry basis weight percentages are listed in order for Cdb,db,Sdb,Odb,Ndb,Adbsi113_e.

b Dry ash-free weight percentages are listed in order for Cdaf,daf,Sdaf,Odaf,Ndafsi114_e.

c Atomic fractions for DAF coal are listed in order for c,h,s,o,nsi115_e and are given in mol/kgdaf.

d Atomic fractions for as-received coal listed in order for c,h,s,o,n,a,msi116_e and given in mol/kgar.

e Atomic fractions per kmol of coal as-received listed in order for c,h,s,o,n,a,msi117_e and given in kmol/kmolcoal.

Table 5.19 gives the calculated stream concentrations, enthalpies, entropies, and chemical exergies for the first phase of the combustion process. For 1 kmol of coal supplied to the combustion process 5.64 kmol of water is circulated in the calandria tubes. Water boils partially up to a quality of 25% in the lower part of the furnace. The combusted volatile matter consisting of CO2, H2O, SO2, and N2 reaches a temperature of 773 K at point A. The bottom ash leaves the furnace at 500 K. Based on the energy balance equation it results that the enthalpy transferred to boiling water in region 0–A of the furnace is 40 MJ/kmol of supplied coal. The generated entropy represents 1.14% of output entropy whereas the destroyed exergy is 4.9% of the exergy at the input. Table 5.20 gives the streams parameters and the balance equations for the fixed carbon combustion zone of the furnace.

Table 5.19

Stream Parameters and Balance Equations for Ignition Zone

StreamCompositionT (K)H (MJ)S (kJ/K)Ex (MJ)
CompoundAmounta
Pulverized coalC0.426298.15− 15.4338.7136.1
H0.512
S0.002
O0.021
N0.005
A0.002
H2O0.032
Fresh airO20.655298.150.06190.004
N22.462
Water inlet (#22)H2O (x = 0) b5.64620.5167.6380.554.6
Fixed carbonC0.098773.00.6971.87740.2
Combusted volatile matterCO21.054773.0− 453.72067101.5
H2O0.719
SO20.006
O20.422
N26.661
AshA0.002500.0− 4.150.4010.195
Boiling water exit (#23)H2O (x = 0.25)5.64620.5222.1468.482.9
Energy balance equationHin = Hout + Qgensteam = − 414.6 MJ; Qgensteam = H23 − H22 = 40.0 MJ
Entropy balance equationSin + Sg = Sout = 2, 191 kJ/K; Sg = 24.9kJ/K = 1.14 % of Sout
Exergy balance equationExin = Exout + Exd = 153.0 MJ; Exd = 7.4 MJ = 4.9 % of Exin

t0100

a Given in kmol component per kmol coal (as-received basis).

b Formation enthalpy, formation entropy, and chemical exergy not included.

Table 5.20

Streams Parameters and Balance Equations for Carbon Combustion Zone

StreamCompositionT (K)H (MJ)S (kJ/K)Ex (MJ)
CompoundAmounta
Boiling water outlet (#24)bH2O (q = 0.35)5.64620.5234.8488.989.5
Combustion gases in BCO21.1521073− 467.92167136.5
H2O0.778
SO20.006
O20.295
N26.661
Energy balance equationHin = Hout + Qgensteam = − 454.6 MJ; Qgensteam = H24 − H23 = 13.3 MJ
Entropy balance equationSin + Sg = Sout = 2, 218 kJ/K; Sg = 26.8kJ/K = 1.23 % of Sout
Exergy balance equationExin = Exout + Exd = 145.5 MJ; Exd = 8.0 MJ = 5.5 % of Exin

t0105

Note: balances are given per unit of kmol of coal. The input streams are outputs of previous unit (Table 5.16).

a kmol component per kmol coal.

b Formation enthalpy, entropy, and chemical exergy not included for water.

From the table it results that in the calandria tubes of region A–B steam augments its vapor quality from 0.25 to 0.35. The heat input to steam is 13.3 MJ/kmol of coal fed; the entropy generation represents 1.23% of entropy in B while exergy destruction is 5.5% of exergy input in A. The vapor quality of steam at its return to the calandria drums in state 25 (Figure 5.26) is 75%.

The flue gas parameters in the non-combustion zone of the furnace comprising states C, D, E, and F (Figure 5.25) can be observed in Table 5.21. From this table it results that the exergy destruction at the superheater is 7.53 MJ/kmol of coal, whereas exergy destruction of the reheater is 2.18MJ/kmol. The exergy destruction of the economizer is reported as 2.38 MJ/kmol of coal.

Table 5.21

Stream Parameters and Irreversibilities for Furnace’s Non-Combustion Zones

StreamT (K)H (MJ)S (kJ/K)Ex (MJ)Irreversibilities
ComponentSg (kJ/K)Exd (MJ)
C644− 504.3193276.3Superheater25.27.53
Reheater7.332.18
D630− 507.9192674.4Calandria boiler0.150.05
E473− 550.1178438.4Economizer8.002.38
Fa473220713.615.4Total40.712.13

t0110

Note: the values of enthalpies, entropies, and exergies are given for 1 kmol of coal combusted.

a Flue gas recirculation ratio: 0.6.

The balance equations are written for all components of the plant. A system of algebraic equations is formed which can be solved for thermodynamic parameters (temperature, pressure, enthalpy, entropy, exergy) at each state point. Values for the isentropic efficiency of the pumps and turbines must be assumed.

The imposed known parameters for modeling are the water condensation temperature at the condenser inlet and outlet, temperature of coal at feed (taken as T0), boiling temperature (T8), and superheating (T9) and reheating (T10) temperatures. Furthermore, the minimum temperature difference for heat exchangers is taken as ~10 K.

With these assumptions the algebraic equation system is closed and can be solved. Once thermodynamic state points are determined the irreversibilities in the form of entropy generation and exergy destruction are determined for each system component.

The system components are grouped into four major subsystems: furnace, turbines, condenser, and feedwater heating system. The results for each thermodynamic state are given in Table 5.22. The exergy destruction for each of the subsystems is shown graphically in Figure 5.30. The maximum exergy destruction is in the furnace, followed by the turbines. The HPT destroys 8.2% exergy, whereas the LPT destroys 31.7%.

Table 5.22

Working Fluid Stream Parameters (Example 5.7)

StreamnPTHSExStreamnPTHSEx
0a1298.150.1010.0026.6-0.082130.4787115.937722,23860.684187
13.844.246303.2869530.25.264140.4787115.9377375311.64321.6
23.84260.6303.2871630.2123.13150.47874.246303.2375312.5744.68
33.84260.636727,20085.542011160.2577260.640212,49632.372867
44.229260.640241,248123.64745171.752551.1428.783,845207.222,223
54.22916,000404.142,750124.26081181.62551.1428.777,564191.620,558
64.22916,000418.747,479135.77383190.1312551.1428.7628115.521665
74.22916,000620.5125,678285.440,948200.1312551.1428.715524.488224.4
84.22916,000620.5196,597399.677,792210.1312260.640215524.511217.5
94.22916,000773.1251,108480.1108,298225.63916,000620.5167,571380.554,597
102.4785530620.5136,008284.551,394235.63916,000620.5222,106468.482,930
112.4785530773.1152,99930960,862245.63916,000620.5234,847488.989,550
123.3624.246303.2136,6274522133255.63916,000620.5238,490494.891,442

t0115

n, number of moles (kmol/kmol coal); P, pressure (kPa); T, temperature (K); H, enthalpy (kJ/kmol coal); S, enthalpy (kJ/K for 1 kmol coal); Ex, exergy (kJ/kmol coal).

a Reference state.

f05-30-9780123838605
Figure 5.30 Exergy destructions and efficiencies of the studied coal-fired Rankine plant.

Figure 5.31 illustrates graphically the amounts of exergy destruction for each of the components of the furnace: calandria boiler (segments 0–A, A–B, C–D), superheater, reheater, and economizer. The subsystem destroys 19 MJ/kmol of coal (or 45.5% of input exergy). Proportionally, the superheater destroys the most exergy among the furnace components at 39.6%, followed by the first calandria segment (0–A) with 30.2% and the reheater with 11.5%.

f05-31-9780123838605
Figure 5.31 Exergy destruction in the furnace.

A similar pie chart is presented in Figure 5.32 to show the exergy destruction of the components of the feedwater heating subsystem. Among the seven components (low-pressure feedwater heater, OFW, throttling valve 14, HPP, high-pressure feedwater heater, throttling valve 2, and LPP) the most exergy destruction (66.6%) happens in the low-pressure feedwater heater.

f05-32-9780123838605
Figure 5.32 Exergy destruction in feedwater heating subsystem.

The energy and exergy efficiencies of the overall system result from Figure 4.24 and are 40.3% (energy) and 58.9% (exergy). Also the specific emissions of CO2 and SO2 can be calculated from Equation (5.15). Accordingly, c=0.426si326_e moles of CO2 are generated for mole of coal combusted; moreover, s=0.002si327_e moles of SO2 are produced for one mole of coal combusted. There is 756 g CO2/kWh net power and 5.04 g SO2/kWh net power. Figure 5.33 indicates the molar composition of flue gas as well as the specific CO2 and SO2 emissions.

f05-33-9780123838605
Figure 5.33 Molar composition of flue gases and specific CO2, SO2 emissions.

5.2.6 Organic Rankine Cycles

Unlike the more common steam Rankine cycles, an organic working fluid is used in the ORCs. Local and small-scale power generation as well as renewable energy systems and low-grade heat recovery systems are the best applications for ORC technology. Some examples of low-temperature heat sources from renewable sources are solar radiation, biomass combustion systems, heat recovery systems from engine exhaust gases, geothermal energy, and ocean thermal energy.

ORC technology has received increased attention in the last decades. The manufacturers of ORC provide a wide range of solutions, based on the temperature level and sources. Table 5.23 gives a list of some existing ORC manufacturers and briefly describes their technology.

Table 5.23

List of some ORC Manufacturers and Technology Descriptions

ManufacturerPower rangeHeat source temperatureTechnology descriptionApplications
ORMAT, US200 KW–72 MW150°–300 °CFluid : n-pentaneGeothermal, WHR, solar
Turboden, Italy200 KW–2 MW100°–300°CFluids: OMTS, Solkatherm
Axial turbines
Geothermal, CHP
Adoratec, Germany350 KW–1600 KW300 °CFluid : OMTSCHP
GMK, Germany50 K–2 MW120°–350 °CFluid: GL160 (GMK proprietary)
3000 rpm multistage axial turbines
Geothermal, CHP, WHR
Koehler-Ziegler, Germany70–200 KW150–270 °CFluid: Hydrocarbons, screw expanderCHP
UTC, US280 KW> 93 °CN/AGeothermal, WHR
CryostarN/A100–400 °CFluids: R245fa, R134a
Radial inflow turbine
WHR
Geothermal
Freepower, UK6 KW–120 KW180–225 °CN/AWHR
Tri-o-gen, NL160 KW> 350 °CTurbo-expanderWHR
50 KW> 93 °CTwin screw expanderWHR
Infinity turbine250 KW> 80 °CFluid: R134a
Radial turbo expander
WHR

t0120

WHR, waste heat recovery; CHP, combined heating and power.

Working fluid characteristics have great influence in determining the cycle configuration. Selection of the working fluid must be done in compliance with the ORC application, which is influenced by the type of heat source, level of temperature, modes of heat transfer, and scale of the application. Pure fluids and fluid mixtures can both be considered. Also, the retrograde characteristic of the working fluid plays a crucial role in determining the cycle configuration. For example, if a retrograde fluid is used, then the expansion of saturated vapor occurs in the superheated region. This is due to the fact that the slope of the vapor saturation curve on a Ts diagram for retrograde fluids is always positive (e.g., toluene). On the other hand, if the fluid is regular then the expansion of saturated vapor occurs in the two-phase region. The slope of the vapor saturation curve of regular fluids is negative in the Ts diagram (e.g., R134a).

One important feature related to organic working fluid refers to the fact that a large pressure ratio over the turbine can be obtained for relatively small pressure differences. For example, with the organic fluid R123 the ORC operates at a boiling pressure of 5.87 bar and a condensation pressure of 0.85 bar. The pressure ratio is thus 6.9 and the pressure difference is 5.02 bar. This point signifies the advantage of using organic fluids as the working fluid in Rankine cycles. The most important fluid properties for organic working fluid selection are:

 Thermodynamic and physical properties such as density, boiling enthalpy, liquid heat capacity, viscosity, thermal conductivity, melting point temperature, normal boiling point, boiling latent heat (must be high), critical temperature, critical pressure, retrograde or regular behavior, and zeotropic or azeotropic characteristic in the case of mixtures.

 Environmental impact and safety factors such as ozone depletion potential (ODP), global warming potential (GWP), atmospheric lifetime (e.g., R11 45 years, R152 0.6 years), toxicity, flammability.

 Availability, stability at high temperature, and cost effectiveness.

Table 5.24 gives the main categories of working fluids for ORCs and the typical applications for each category. Some proposed working fluids are propylene, propane, R227ea, R236fa, RC318, cyclohexane, R245fa, R141b, R365mfc, R11, R113, R114, toluene, fluorinol (CF3CH2OH), R41, R123, R245ca, isobutene, propane–ethane, and siloxanes (MM, MD4M, D4, D5). Table 5.24 lists the working fluids in order of their molecular masses and gives some main parameters, namely normal boiling point, critical pressure and temperature, ODP, and GWP.

Table 5.24

Classes of Working Fluids for ORC and Typical ORC Applications

Class of working fluidCondensationBoilingApplications
Alkanes, fluorinated alkanes, ethers and fluorinated ethers30 °C100 °CGeothermal
R123, iso-pentane, HFE7100, Benzene, Toluene, p-xylene30 °C150–200 °CWaste heat recovery
R123, isopentane, R245ca, R245fa, butane, isobutene, and R-152a55 °C60–150 °CWaste heat recovery; cascading with internal combustion engines
Butyl-benzene, propyl-benzene, ethyl-benzene, toluene90°C250–350 °CBiomass combustion, combined heating and power; concentrated solar power
R245fa, R123, R134a, n-pentane35 °C60–100 °CWaste heat recovery
R123, n-pentane, PF505090 °C70–90 °CGeothermal
R227ea, propylene, RC318, R236fa, isobutane, R245fa25 °C80–120 °CGeothermal

t0125

Both turbo-expanders and positive displacement expanders are considered prime movers for ORCs. The choice depends on the scale of the application, the working fluid, pressure differentials, and pressure and volume ratios. For large-scale applications turbo-expanders will be the preferred choice because of their higher efficiency. At lower-scale applications, when pressure drop across the prime mover must be high, the preferred choice is a positive displacement expander because it has contact seals which reduce the bypass leakages. When pressure difference per stage is high, turbo-expanders (turbines) have bypass leakage flow at blade tips which is accentuated for small-scale applications.

There are many configurations of the ORC which differ from the typical reheat-regenerative steam cycle configuration that normally uses steam extraction. For example, in ORC with retrograde fluid it is possible to use a counter-current heat exchanger as a regenerator. In addition, ORC with retrograde fluids can operate without superheating, which makes them better suited for applications that involve heat recovery or solar energy. Other configurations are the reheat-regeneration cycle and supercritical cycle.

An ORC of simple configuration is shown in Figure 5.34. The cycle operates with the organic fluid R134a and is heated with a hot fluid. This cycle configuration can be applied for waste heat recovery when the heat source is low grade. Notice that superheating is not required because the moisture at the end of the expansion process is very low when saturated vapors are expanded (Table 5.25).

f05-34-9780123838605
Figure 5.34 Simple four-component ORC with R134a for low-grade waste heat recovery.

Table 5.25

Working Fluids Commonly Used in ORC

FluidM (kg/kmol)NBP (°C)Tc (°C)Pc (MPa)ODPGWP
Methanol32.0464.4240.28.104N/AN/A
R29044.1− 42.196.684.2470~ 20
Ethanol46.0778.4240.86.148N/AN/A
R3252.02− 51.778.115.7840675
R600a58.12− 11.71353.6470~ 20
R60058.12− 0.51523.7960~ 20
R152a66.05− 24113.34.520124
R60172.1536.1196.53.3640~ 20
Cyclohexane84.1680.7280.54.075N/AN/A
R407c86.2− 43.686.794.59701800
Toluene92.14110.6318.64.126N/AN/A
R50099.3− 33.6105.54.4550.7388100
R134a102.03− 26.11014.05901430
R141b116.9532204.24.2490.12725
R12120.91− 29.81124.114110,890
R123152.9327.8183.73.6680.0277
Biphenyl154.22555473.25N/AN/A
R114170.923.6145.73.289110,040
R113187.3847.6214.13.43916130
RC318200.03-6115.22.778010,250

t0130

The regenerative schemes for ORC do not require (in general) vapor extraction (as in the case of steam cycles). Rather an additional heat exchanger is inserted between the turbine exit and the pump exit, which allows for a transfer of heat between the hotter working fluid at the turbine exhaust and the colder liquid at the pump discharge. The configuration of a regenerative ORC is presented in Figure 5.35. An example of ORC with regeneration that uses R134a as working fluid is presented in Figure 5.36. In this cycle, boiling occurs at 353 K; it starts with saturated liquid in state 4a and reaches saturated vapor in state 4b (pressure drop is neglected). In the condenser, there is a de-superheating process according to line 6–1a, where 1a represents saturated vapor (see Figure 5.36). Further, the condensation process continues until reaching saturated liquid in state 1.

f05-35-9780123838605
Figure 5.35 Regenerative ORC configuration.
f05-36-9780123838605
Figure 5.36 Regenerative ORC with R134a for low-grade heat sources.

The exergy destruction in the regenerator is accounted for with the help of regenerator effectiveness. This parameter is defined based on the sketch in Figure 5.37, which illustrates a regenerator in the form of a heat exchanger that facilitates a heat transfer process between a hotter stream and a colder stream. Both streams exchange sensible heat. The effectiveness is defined by the ratio between the heat transfer rate that ideally can be extracted from the hotter stream (Q˙idealsi328_e) and the heat which is actually received by the colder stream (Q˙=m˙colderh4-h3)si329_e. Note that the ideal, maximum heat rate that the hotter fluid can transfer occurs when T2 = T3 (see Figure 5.37); therefore the ideal maximum heat transfer in the regenerator occurs when h2 = h2(T3) and is given by Q˙ideal=m˙hotter(h1-h2(T3))si330_e. According to this definition, the regenerator effectiveness is

ɛreg=Q˙Q˙ideal=m˙colderh4-h3m˙hotterh1-h2T3

si331_e  (5.23)

f05-37-9780123838605
Figure 5.37 Sketch for the definition of regenerator effectiveness.

An interesting type of regenerative ORC is that using retrograde working fluids where saturated vapor is fed into the turbine while the expanded fluid is in superheated state. An example of such a cycle is shown in Figure 5.38, with toluene as the working fluid. As seen, the boiling temperature is 573 K and the boiling pressure is 1665 kPa. For a condensation at 303 K the saturation pressure is 2.911 kP (vacuum). Therefore the pressure ratio over the turbine is 395; at the same time the volumetric expansion ratio is 285. Figure 5.38 is suggested the temperature profile at the source when a hot fluid which exchanges sensible heat is used as the heat source (e.g., hot combustion gases, high-temperature geothermal brine, etc.).

f05-38-9780123838605
Figure 5.38 ORC with a retrograde working fluid (toluene) and expansion of saturated vapor.

An emerging type of ORC is the supercritical cycle. This cycle works generally based on a scheme with regeneration such as that presented in Figure 5.35. In supercritical cycles the working fluid is pressurized to supercritical pressure and heated. As the pressure is supercritical, there is no boiling during heating. Rather, when the temperature becomes superior to critical point the working fluid enters into a supercritical state where the thermodynamic properties are neither of a liquid nor of a gas but a combination of these. Given that there is no boiling, the typical pinch point problem that occurs in vapor generators for Rankine cycles is minimized. Therefore, a gain on exergy efficiency can be obtained with respect to cycles with boiling. However, due to the high pressures associated with supercritical cycles, one must exercise care when selecting a working fluid.

An example of a supercritical ORC operating with R404A working fluid (which is a blend of refrigerants) is shown in Figure 5.39. The high pressure in this cycle is 44.82 bar while the low pressure is 14.28 bar, making a pressure ratio of 3.1, which is applicable to positive displacement expanders. The temperature profiles match very well, proving that there is reduced exergy destruction in a supercritical ORC. In the regenerative heat exchanger a pinch point appears due to the nonlinear temperature profile of the supercritical fluid.

f05-39-9780123838605
Figure 5.39 Supercritical-regenerative ORC with R404a (refrigerants blend).

Example 5.8

In this example the thermodynamic analysis of a regenerative cycle with the retrograde fluid toluene is demonstrated. The cycle configuration is according to Figure 5.35. Assume that the boiling temperature is 525 K, condensation is at 305 K cooled with water, the isentropic efficiency of the pump is 0.8, and isentropic efficiency of the turbine is 0.9. The heat source is a hot flue gas of 650 K that is assumed as ideal gas air. The minimum temperature difference in the heat exchangers is 10 K and the log mean temperature difference (LMTD) at the condenser is 18 K. We will determine the efficiencies, the exergy destructions, and the effectiveness of the regenerator and the mass flow rates for 100 kW power generated.

The balance equations in conjunction with the algebraic equations for pump and turbine isentropic efficiency and the temperature assumptions form a closed system that can be solved to determine the thermodynamic parameters for each state point. Table 5.26 gives the assumptions and balance equations for each of the cycle components.

Table 5.26

Balance Equations for Example 5.8

ComponentProcessAssumptionsBalance equations
Pump1–2h2s − h1 = ηS,p(h2 − h1) with ηS,p = 0.8 adiabatic; x1 = 0MBE: m˙1=m˙2=m˙si118_e
EBE: m˙1h1+Wp˙=m˙2h2si119_e
EnBE: m˙1s1+S˙g=m˙2s2si120_e
ExBE: m˙1ex1+Wp˙=m˙2ex2si121_e
Regenerator2–3 and 5–6T6 = T2 − 10

ɛreg=h3h2h5hP5,T2si122_e, adiabatic
MBE: m˙2=m˙3=m˙5=m˙6=m˙si123_e
EBE: m˙2h2+m˙5h5=m˙3h3+m˙6h6si124_e
EnBE: m˙2s2+m˙5s5+S˙g=m˙3s3+m˙6s6si125_e
ExBE: m˙2ex2+m˙5ex5=m˙3ex3+m˙6ex6+E˙xdsi126_e
Vapor generator3–4Tso,1 = T4 + 10
Tso,pinch = T4a + 10, adiabatic, isobaric
MBE: m˙3=m˙4=m˙;m˙so,1=m˙so,2=m˙sosi127_e
EBE: m˙3h3+m˙so,1hso,1=m˙4h4+m˙so,2hso,2si128_e
EnBE: m˙3s3+m˙so,1sso,1+S˙g=m˙4s4+m˙so,2sso,2si129_e
ExBE: m˙3ex3+m˙so,1exso,1=m˙4ex4+m˙so,2exso,2+E˙xdsi130_e
Turbine4–5h4 − h5 = ηS,t(h4 − h5s)
with ηS,t = 0.9, adiabatic
MBE: m˙4=m˙5=m˙si131_e
EBE: m˙4h4=m˙5h5+Wt˙si132_e
EnBE: m˙4s4+S˙g=m˙5s5si133_e
ExBE: m˙4ex4=m˙5ex5+Wt˙+E˙xdsi134_e
Condenser6–1ΔTpinch = T1a − Tsi = 10 K
LMTDsi = 10 K, adiabatic, isobaric
MBE: m˙6=m˙1=m˙;m˙si,1=m˙si,2=m˙sisi135_e
EBE: m˙6h6+m˙si,1hsi,1=m˙1h1+m˙si,2hsi,2si136_e
EnBE: m˙6s6+m˙si,1ssi,1+S˙g=m˙1s1+m˙si,2ssi,2si137_e
ExBE: m˙6ex6+m˙si,1exsi,1=m˙1ex1+m˙si,2exsi,2+E˙xdsi138_e

t0135

The chemical exergies do not need to be considered because there is no chemical reaction. For the thermomechanical exergy calculation the standard reference state of T0 = 298.15 K and P0 = 101.325 kPa is considered. For toluene, which is always enclosed inside the heat engine circuits, the reference state is saturated liquid at T0.

A pinch point calculation procedure must be applied for both the vapor generator and the condenser. In this respect, the requirement of a 10 K-minimum temperature difference is called. Hence, the temperature of the gas at source side must be superior by 10 K with respect to the working fluid temperature at state 4a, which corresponds to the pinch point. In addition we assume that the temperature difference between the source inlet temperature and the working fluid temperature at the turbine entrance is 10 K. Regarding the condenser, a temperature difference between the saturated vapor (state 1a) and the water coolant of 10 K is imposed (this corresponds to the pinch point). Further, the temperature of water at inlet must be iterated until LMTD at the condenser becomes 18 K. The LMTD must be calculated with ΔT1 = T1 − Tsi,1 where Tsi,1 is the water temperature at sink inlet and ΔT2 = T6 − Tsi,2 where Tsi,2 is the water temperature at sink outlet. The known equation for LMTD is: LMTD = (ΔTmax − ΔTmin)/ln(ΔTmax/ΔTmin) where ΔTmin = min(ΔT1, ΔT2) and ΔTmax = max(ΔT1, ΔT2).

The temperatures, pressures, and enthalpy, entropy, and exergy rates for each thermodynamic state of the cycle are given in Table 5.27. In the same table the rates of heat and work transfer and the rate of exergy destruction for each cycle component are also given. The energy efficiency of the cycle is 38% while the exergy efficiency of 50%. Based on 100 kW net power generation the mass flow rate of the working fluid must be 0.3786 kg/s.

Table 5.27

Thermodynamic State Points, Components, and Cycle Parameters for Example 5.8

StateT (K)P (kPa)H˙si139_e (kW)S˙si140_e (kW/K)Ex˙si141_e (kW)ComponentQ˙si142_e(kW)W˙si143_e (kW)E˙xdsi144_e (kW)
13055.362− 56.78− 0.1650.05225PumpN/A0.960.188
2305.71710− 55.82− 0.16440.8287Regenerator119.7N/A6.613
3459.9171063.840.149426.92Vapor generator263.3N/A8.668
4a5251710126.40.276551.63TurbineN/A100.966.615
4b5251710216.20.447690.4Condenser163.3N/A7.187
46501710327.20.6371144.9Cycle parameters
5513.35.362226.20.659337.29Energy efficiencyη = 38%
6315.75.362106.60.37033.804Exergy efficiencyψ = 50%
1a3055.362101.70.35264.182Mass flow rateWorking fluidm˙=0.3786si145_e kg/s
1so660101.3102511.73198Source fluidm˙so=1.528si146_e kg/s
2so535101.3824.611.3997.69Sink fluidm˙si=6.159si147_e kg/s
1si288.9101.3406.11.443− 24.24TurbinePressure ratioPR = 319
2si295101.3564.61.986− 27.63Expansion ratioER = 285

t0140

In Figure 5.40a the thermodynamic cycle is illustrated in the Ts diagram. On the same diagram the temperature profiles of the heat source and heat sink fluids are suggested. This illustrates the irreversibility due to the finite temperature difference for heat transfer at heat source and heat sink heat exchangers. The exergy destructions by the cycle components are represented in the chart in Figure 5.40b. As seen, the exergy destructions are well balanced among the cycle components, except for the pump, which destroys negligible exergy.

f05-40-9780123838605
Figure 5.40 ORC with toluene for Example 5.8; (a) thermodynamic cycle, (b) exergy destruction percentages.
..................Content has been hidden....................

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