Sie sind auf Seite 1von 10

Proceedings of ECOS 2009 Copyright 2009 by ABCM

22 International Conference on Efficiency, Cost, Optimization Simulation and Environmental Impact of Energy Systems August 31 September 3, 2009, Foz do Iguau, Paran, Brazil

nd

ON THE TREATMENT OF DISSIPATIVE COMPONETS AND RESIDUES IN THERMOECONOMIC MODELING


Jos Joaquim Conceio Soares Santos, jjsantos@unifei.edu.br Marco Antnio Rosa do Nascimento, marcoantonio@unifei.edu.br Electo Eduardo Silva Lora, electo@unifei.edu.br Jos Carlos Escobar Palacio, jocescobar@gmail.com Julio Augusto Mendes da Silva, juliomendes@unifei.edu.br
Federal University of Itajub UNIFEI Excellence Group in Thermal Power and Distributed Generation NEST Av BPS 1303, CP 50, CEP 37500-903, Pinheirinho, Itajub-MG Brazil

Abstract: One of the thermoeconomic methodologies challenges is to define the productive structure of thermal systems that allows to allocate rationally the cost of the residues and the cost of the dissipative components to the final products. The allocation of the residues and dissipative components is a very important subject regarding the inclusion of the environmental costs in thermoeconomics in the near future. Although there has been an advance in the development of thermoeconomic criteria for residues and dissipative components allocation, this problem is still open. The way in which we define the productive structure is a key point of the thermoeconomic modeling. Some approaches propose a productive structure using only exergy flows (E Model). Others propose a productive structures that include the negentropy as a fictitious flow, joined up with exergy flows (E&S Model). A more recent approach proposes the use of the physical exergy disaggregated into enthalpy and syntropy (negative entropy), combined with the chemical exergy (H&S Model). This work compares these three kinds of productive structures by applying them for thermoeconomic modeling in a combined cycle power plant. The goal is to calculate the exergetic unit cost of the internal flows and product. These applications show that the productive structure that uses only total exergy does not allow to show explicitly the allocation of the residues nor allows to isolate the dissipative components. This is possible by including negentropy flows in the productive structure. But the use of negentropy flow leads to some ambiguousnesses because the product of some components of the productive structure is greater than their fuels. This approach also needs special considerations to formulate the auxiliary equations. These ambiguousnesses and special considerations to formulate the auxiliary equations can be avoided by disaggregating the physical exergy into enthalpy and syntropy. Keywords: Thermoeconomic Modeling, Dissipative Component, Residue, Syntropy, Negentropy 1. INTRODUCTION Unavoidably, in any productive process, the achievement of functional products is inseparable from the generation of residues and waste disposals. Therefore, in the same way as there is a process of cost formation of the functional products, there also exists a cost formation process of the residues. These residues are disposed off into the environment and if they are wanted to be eliminated, then additional resources are required to diminish their intensive potential either by cooling, expansion or dispersion. The conventional methods of thermoeconomic cost accounting, generally do not consider the in-depth analysis of the cost allocation of residues. They have placed emphasis on the production process traditionally, without bearing in mind the cost of residues. The cost formation process of the residues becomes essential when it comes to eliminating the residues by means of some sort of pollution control system, and it is required to allocate rationally their costs to the functional system products. Therefore, the need is evident for either developing new techniques or extending the existing ones to include both the cost allocation of residues and the analysis of their formation process. The residue cost allocation is a complex problem since it depends on the nature of such flows and how they have been formed. Every residue has a cost formation process that must be identified to calculate the cost of all functional products correctly. The cost of the exergy contained in the residues must be allocated to the productive units that have generated them, and thus to the final products, according to the productive structure of the system. This allocation can be made in several ways, depending on the type and nature of the residue, as mentioned above. However, there is not a general criterion to define the residue cost distribution ratios (Torres et al., 2008). In any energy system, in the same manner that there are productive components, there also exist dissipative components, whose purpose is to partially or totally eliminate the undesirable flows. Their utility lies in interacting with other components, which in some cases allows the system to have a higher production or a better efficiency. The exergy of residues eliminated in a dissipative equipment and its formation cost is as objective as the exergy and the cost of any other functional product. An energy system has a defined productive structure, but also a dissipative one and both structures are not independent. The choice of the best residue distribution among possible alternatives is still an open research line. Although there has been an advance in the development of criteria for the cost allocation of residues, this problem is still open. In this regard, the definition of an appropriate productive structure will help to identify where the
63

residues have been formed and to allocate their costs (Torres et al., 2008). The way in which we define the productive structure is a key point of the thermoeconomic modeling (Lozano and Valero, 1993). Thus, this paper compares three kinds of productive structures by applying them for thermoeconomic modeling in a combined cycle power plant. In a combined cycle power plant, the residues are the exhaust gases released to the atmosphere through the chimney of the recovery boiler, and the dissipative component is the condenser. The goal is to allocate the external fuel exergy to determine the exergetic unit cost of the internal flows and final product. The first methodology (here called E Model) defines the productive structure by using only total exergy flows. The second (here called E&S Model) uses total exergy and negentropy flows, i. e., the negentropy is considered as a fictitious flow. The third methodology (H&S Model) defines the productive structure by disaggregating the exergy flow into chemical, enthalpic and syntropic components, i. e., the syntropy (negative entropy) is a physical exergy component flow. 2. PHYSICAL STRUCTURE AND THERMODYNAMIC MODEL Figure 1 represents the physical structure of the combined cycle power plant, which is defined as having eight units or subsystems: air compressor (AC), combustion chamber (CC), gas turbine (GT), electric generator (EG), recovery boiler (RB), feeding pump and its motor (P), steam turbine (ST) and condenser and condensing supply works (C).

Figure 1: Physical Structure of the Combined Cycle Power Plant The streams of the physical structure of the combined cycle power plant (Figure 1) are: air, gases, water, steam, liquid vapor mixture, mechanical and electric power and natural gas. The description of the streams that represent the working fluids and their main parameters (mass flow, pressure and temperature) are presented in Table 1. Table 1: Main Parameters of the Main Physical Flows of the Combined Cycle Power Plant i 1 2 3 4 5 6 7 8 9 PHYSICAL FLOW Description Air Air Gases Gases Gases Water Water Steam Moisture (x = 0,89) m [kg/s] 310.00 310.00 314.20 314.20 314.20 29.08 29.08 29.08 29.08 p [kPa] 101.32 911.90 902.80 104.33 101.32 6.50 4,080.00 4,000.00 6.50 T [C] 25.00 331.78 870.00 443.05 180.64 37.63 38.35 417.00 37.63

The mechanical net power (b) of the Brayton cycle is 58,280.05 kW and the compressor power (a) is 98,893.09 kW. The steam turbine produces 27,903.64 kW of mechanical power (c). The fuel is natural gas (ng), whose consumption in exergetic basis is 209,697.64 kW. The gross electric power (gp) produced by the electric generator is 84,415.00 kW. The electric power consumed by the feeding pump motor (d) and by the condensing supply works (e) are 205.00 kW and 210.00 kW, respectively. To model the combined cycle power plant represented in the Figure 1, the thermodynamic
64

Proceedings of ECOS 2009 Copyright 2009 by ABCM

22 International Conference on Efficiency, Cost, Optimization Simulation and Environmental Impact of Energy Systems August 31 September 3, 2009, Foz do Iguau, Paran, Brazil

nd

model considers complete combustion with excess of air and it also considers that the air and the combustion gases are mixtures of ideal gases. The molar compositions of the air and the combustion gases streams are in Table 2. Table 2: Molar Composition of Air and Gases Streams present in the Physical Structure of the Combined Cycle Plant n 1 2 3 4 5 ELEMENT Description Oxygen Carbon Dioxide Water Vapor Nitrogen Argon PERCENTAGE [%] Symbol O2 CO2 H2O N2 Ar Air 20.56 0.03 1.88 76.61 0.92 Gases 15.56 2.36 6.26 74.92 0.90

The thermodynamic model considers that the specific heat (Cp) of the elements that composes the streams of air and gases varies with their temperature, according to the polynomial equation and the respective coefficients in the Table 3. Table 3: Coefficients for the Specific Heat Polynomial Equation of some Ideal Gases (Lozano and Valero, 1986) ELEMENTS Description Oxygen Carbon Dioxide Water Vapor Nitrogen Argon**
Cp = A + B T + C T 2 + D T 3 [kcal/kmol.K]

n 1 2 3 4 5**

Symbol O2 CO2 H2O N2 Ar**

A 6.085 5.316 7.7 6.903 4.964**

B 10 2 0.3631 1.4285 0.04594 -0.03753 0.00

C 105 -0.1709 -0.8362 0.2521 0.193 0.00

D 109 0.3133 1.784 -0.8587 -0.6861 0.00

OBS: ** (Verda et al., 2004)

In order to carry out the thermoeconomic modeling aimed by this paper, some thermodynamic magnitudes, such as exergy (E), negentropy or syntropy (S) and enthalpy (H), need to be calculated for each of the working fluid streams (i). For the physical flows representing water, steam and liquid vapor mixture (i = 6, 7, 8 and 9), the exergy (Ei), the negentropy or the syntropy (Si) and the enthalpy (Hi) should be calculated by using the Eqs. (1), (2) and (3). The variables that define these three equations are: mass flow (mi), enthalpy (hi) and entropy (si) of the stream i; and enthalpy (h0), entropy (s0) and temperature (T0 = 25oC) of water at the conditions of thermodynamic environment 0.
E i = mi [(hi h0 ) T0 ( s i s 0 )] S i = m i T0 ( s i s 0 ) H i = m i ( hi h 0 )

(1) (2) (3)

The streams representing air and gas (i = 1, 2, 3, 4 and 5) are mixtures of ideal gases (n = N2, CO2, O2, H2O and Ar). Once their chemical composition is different from that of the air, besides the exergy (Ei), the negentropy or syntropy (Si) and the enthalpy (Hi), the chemical exergy (CHi) needs to be calculated, as show Eqs. (4), (5), (6) and (7), respectively.
Ei = mi

y n ,i

y n ,i y n,i R T0 ln y Mn n ,o

Ti + y n,i Cp n 1 T0 T i T0

p dT + R T0 ln i p 0

(4)

Si =

Ti p i m i T0 Cp n y n ,i dT R ln p Ti y n ,i M n 0 T0

(5)

Hi =

y n ,i M n
y n ,i
mi

mi

y n,i Cp n dT
T0

Ti

(6)

CH i =

y n ,i y n,i R T0 ln y n ,o Mn

(7)

65

The new variable that define these four equations are: molar fraction of each element of the stream (yn,i) or of the air (yn,0), molecular weigh of each element (Mn), universal gas constant (R), specific heat of each element (Cpn), temperature of the stream (Ti) and pressure of the stream (pi) or of the thermodynamic environment (p0 = 101.32 kPa).
3. THERMOECONOMIC MODELING

The thermoeconomic model is a set of equations which describes the cost formation process of the system. But, the physical model is not enough to identify the formation process of the residues. Perhaps the fundamental limitation of the theory of exergetic cost, as it was originally formulated, consisted of defining the productive structure in relation to the same flows and component present in the physical structure. The resulting difficulties lie mainly in the adequate treatment of the dissipative units and of the residues (Lozano and Valero 1993). To carry out a thermoeconomic analysis of a system, it is convenient to make up a thermoeconomic model, which define the productive propose of the subsystems (products and fuels), as well as the distribution of the external resources and internal product throughout the system. It could be represented by means of the productive or functional diagram. The only limitation with must be imposed it that it be possible to evaluate all the flows of the productive structure in relation to the state of the plant as defined by the physical structure (Lozano and Valero 1993). As mentioned above, in order to define the productive structure, this paper considers three different methodologies, i. e., three ways to define the internal flows. The first uses only the exergy (E Model). The second use the exergy and the negentropy (E&S Model). The third uses the enthalpy, the syntropy and the chemical exergy (H&S Model).
3.1 E Model: total exergy only

Figure 2 shows the productive structure of the combined cycle power plant represented by means of the functional diagram, which represents all the cost formation process in the system. The only external exergy resource consumed by the system is natural gas (Qng) and the only functional product is the electric net power (Pnp).

Figure 2: Productive Diagram of the Combined Cycle Power Plant according to the E Model In this productive diagram (Fig. 2), the rectangles are the real units that represent the actual components of the system. The rhombus and the circles are fictitious units called junction (JEb, JEr and JPm) and bifurcations (BEb, BPm and BPe), respectively. Each productive units of Fig. 2 has inlet and outlet arrows, that represent its fuel (or resource) and
66

Proceedings of ECOS 2009 Copyright 2009 by ABCM

22 International Conference on Efficiency, Cost, Optimization Simulation and Environmental Impact of Energy Systems August 31 September 3, 2009, Foz do Iguau, Paran, Brazil

nd

products, respectively. All of the flows of the productive structure are exergies that represents the mechanical power (Pa, Pb and Pc), the electric power (Pd, Pe, Pgp and Pnp), the external fuel consumption (Qng), or the exergy added to and removed from the working fluid (Ej:k). Each productive flow is defined based on physical flows. The productive flows representing the exergy added to and removed from the working fluid (Ej:k) are always exergy variations between two physical flows (Ej and Ek), as shows Eq. (8). The thermodynamic model to determine the exergies of the working fluid streams (physical flows) was presented above through the Eqs. (1) and (4). The equipments that add exergy to the working fluid produce exergy and, the equipments that remove exergy from the working fluid consume exergy.
E j:k = E j E k

(8)

The E Model uses only exergy to describe the fuels and the product of the subsystems. Once the condenser does not have a product that can be measured in exergetic terms, the steam turbine (ST) and the condenser (C) have been analysed as a single unit (ST-C), once that the authors (Arena and Borchiellini, 1999) generally assume that the condenser increase the steam turbine capacity to produce mechanical work. The mathematical model for external fuel exergy allocation is obtained by formulating cost equation balance in each unit (real and fictitious) of the productive structure, as shows Eq. (9), where c is the exergetic unit cost (unknown variable) of each productive flow represented generically by Y. The variable Y can be mechanical power (Pa, Pb and Pc), electric power (Pd, Pe, Pgp and Pnp) or exergy added to and removed from the working fluid (Ej:k).

(c Y ) = 0

(9)

In order to formulate the cost equation balance in each productive unit or subsystem, conventionally, the inlet flows (fuels) assume negative value and the outlet flows (products) assume positive value. The number of flows is always greater than the number of units. In order to determinate the set of cost equations, some auxiliary equations are necessary. The auxiliary equations attribute the same exergetic unit cost to all of the flows leaving the same bifurcation. This is a common rules used (or accepted) by all thermoeconomic practitioners to formulate the auxiliary equations. According to this approach (E Model), the cost of the residues (the exergetic unit cost of stream 5, which represents the exhaust gases) is implicitly allocated to the gas turbine (GT) and recovery boiler (RB) proportionally to the respective fuel value of these two productive units (E3:4 and E4:5). Thus, the residue cost is allocated to the two products of the Brayton cycle: the power produced by the gas turbine (Pb) and the exergy produced by the recovery boiler (E8:7). The solution of the set of cost equations allows the attainment of the exergetic unit cost of each internal flow and final product. The same exergetic unit cost of the Brayton cycle products obtained by the E Model, can also be reached by allocating the residues to the combustion chamber (where it is originated), as proposed by Torres and Valero (2000) or by Lazzaretto and Tsatsaronis (2006). According to the original formulation of the Exergetic Cost Theory (Torres et al., 1996), the residue cost is allocated entirely to the recovery boiler only, which overcharges the cost of the exergy produced by the recovery boiler to the detriment of the cost of mechanical power produced by the gas turbine.
3.2 E&S Model: total exergy and negentropy

Figure 3 shows the productive diagram of the combined cycle power according to the E&S Model, which uses exergy flows (Ej:k) and negentropy flows (Sj:k) to define the fuels and the products of its subsystems. The negentropy concept has been introduced in thermoeconomics to define the product of the condenser (Frangopoulos, 1987). In a Rankine cycle, the pump and the boiler produce exergy. The turbine consumes part of this produced exergy to generate work. The operation of these units increases the entropy of the working fluid. This increase of entropy must be rejected to the environment through the condenser. In other words, the condenser provides the necessary negentropy for the correct cyclical operation of the system (Lozano and Valero, 1993). According to Frangopoulos (1987), the condenser is supplying the system with the negative of entropy (negentropy, as introduced by Brillouin and used by Smith to quantify the function of the condenser). Thus, the costs associated with the condenser are distributed between all the productive units that increase the working fluid entropy, instead of being charged only to the steam turbine. In the case of a gas turbine cogeneration system, the residual exergy of the gases which abandon the plant through the chimney of the recovery boiler can be explained by the increase of the working fluid entropy in the compressor, in the combustor and in the turbine. The recovery boiler has a negative contribution (Lozano and Valero, 1993). Thus, the recovery boiler produces negentropy. Lozano and Valero (1993) consider that the other part of negentropy is produced by the chimney, an imaginary dissipative unit (here called environment), where residual exergy of the gases is charged. This negentropy, plus that produced by the recovery boiler are given to the units that increase the working fluid entropy. Alves and Nebra (2006) call this imaginary dissipative unit stack (instead of chimney or environment). Comparing the productive diagram defined with exergy flow only (Fig. 2) to this one that include negentropy flow (Fig. 3), some new flows and units appear: negentropy flows (Sj:k), bifurcation of negentropy for the Rankine cycle (BSr), bifurcation (BSb) and junction (JSb) of negentropy for the Brayton cycle, condenser (C) separated from the steam turbine and the imaginary dissipative unit called environment (E) where a new exergy flow (Ej:k) representing the residues is
67

charged. Once the negentropy flows are included in the productive diagram, the real productive units have small junctions to indicate that they have two types of fuel, and the recovery boiler have a small bifurcation because it has two outlet flows (negentropy and exergy). The negentropy flows (Sj:k) of the productive diagram are defined based on physical flows (Sj and Sk), as shows Eq. (10). The physical flows negentropy (Si) is calculated by the Eqs. (2) and (5).
S j:k = S j S k

(10)

The mathematical model is obtained by applying the Eq. (9) of cost balance to each unit of the productive diagram, as described in Section 3.1. In this case, we also have the auxiliary equations that attribute the same exergetic unit cost to all of the flows leaving the same bifurcation (BEb, BPm, BSb, BPe, BEr and BSr). As mentioned above, this is the common rules used (or accepted) by all thermoeconomic practitioners to formulate the auxiliary equations. But when exergy and negentropy (E&S Model) is used to define the productive structure, another auxiliary equation is necessary because the recovery boiler (RB) have two outlet flows. This kind of productive structure was originally proposed and utilized by optimization approaches: Thermoeconomic Functional Analysis (Frangopoulos, 1987) and Engineering Functional Analysis (von Spakovsky,, 1994). Another thermoeconomic approach, named Structural Theory of Thermoeconomics, modified this technique in order to be applied for cost accounting (Lozano and Valero, 1993). The Structural Theory of Thermoeconomics considers that each productive unit can have only one product and the main function of the recovery boiler is to produce exergy. Thus, this approach considers that the negentropy flows exiting the recovery boiler is a by-product, once that its product is the exergy flow. This by-product assumes the same unit cost as the product of the environment (E) that is the only device that produces only negentropy flow. This auxiliary equation is formulated based on this here called Special Consideration, used by the Structural Theory of Thermoeconomic only.

Figure 3: Productive Diagram of the Combined Cycle Power Plant according to the E&S Model This here called Special Consideration has been also used by the Structural Theory of Thermoeconomic to attribute cost to the thermal exergy component produced by the compressor of the Brayton cycle, because this approach considers that the product of the compressor is the mechanical exergy, and consequently, the thermal exergy is the byproduct, which is costed at the same unit cost as the product of the combustor that is the only equipment that produces only thermal exergy in the cycle (Lozano and Valero, 1993). Many thermoeconomic practitioners (Tsatsaronis and Pisa, 1994: Torres et al., 1996) does not accept (or does not use) this Special Consideration, because they attribute

68

Proceedings of ECOS 2009 Copyright 2009 by ABCM

22 International Conference on Efficiency, Cost, Optimization Simulation and Environmental Impact of Energy Systems August 31 September 3, 2009, Foz do Iguau, Paran, Brazil

nd

the same unit cost to the two physical exergy component flows produced by the compressor of the Brayton cycle, which is in accordance with one of the proposition of the Exergetic Cost Theory developed by Valero et al., (1994). Once the formation of residue is justified by the increase of the working fluid entropy, the residue cost is calculated and charged explicitly to the units responsible for the increase of the working fluid entropy of the Brayton cycle (compressor, combustor and turbine), proportionally to the work fluid entropy increased by each of them. In the same way, the cost associated to the condenser is charged to the units responsible for the increase of the working fluid entropy of the Rankine cycle (pump, boiler and turbine), proportionally to the working fluid entropy increased by each of them. In other words, this treatment redistributes the cost of the residues and dissipative component by warding credit when the process decreases the flow entropy, and penalizing it when it increases the entropy.
3.3 H&S Model: enthalpy and syntropy

Figure 4 shows the productive diagram of the combined cycle power according to the H&S Model, which uses enthalpy flows (Hj:k), syntropy flows (Sj:k) and chemical exergy flows (CHj:k) to define the fuels and the products. The terms negentropy and syntropy mean essentially the same thing. However, negentropy is a fictitious flow and syntropy is a physical exergy component. Alves and Nebra (2003) stated that negentropy is a physical exergy component. Other authors (Tsatsaronis and Pisa, 1994; Frangopoulos, 1994) define productive structure by disaggregating the physical exergy into thermal and mechanical components. But the H&S Model was the first that proposes the productive diagram by disaggregating the physical exergy into enthalpic and syntropic (negentropic) components (Santos et al., 2006).

Figure 4: Productive Diagram of the Combined Cycle Power Plant according to the H&S Model By considering the syntropy as an exergy component flow, the H&S Model takes all the advantages due to the use of negentropy in the productive diagram for residue and dissipative component costs allocation, i. e., it wards credit to the processes that decrease the working fluid entropy, and penalizes the ones that increase the working fluid entropy. But, this approach (H&S Model) believes that by using the negentropy joined up with the exergy (E&S Model), some productive subsystems (such as the turbines) are penalized twice due to the increase of the work fluid entropy, while others (such as the recovery boiler, the condenser and the environment) are awarded twice due to the reduction of the work fluid entropy, because the exergy flow already contains the term (m.T0.s) that define the negentropy/syntropy and, consequently, there are productive units (such as the condenser and the environment) where their fuels are less than their products, which from the thermodynamic point of view can be interpreted as an ambiguousness. This

69

ambiguousness is avoided by using the syntropy as an exergy component joined up with the enthalpic and the chemical component. This fact can be proved by analyzing the Eqs. (1)-(7) and the value of the productive flows in the Table 4. Comparing the productive diagram defined by the E&S Model (Fig. 3) to this one defined by H&S Model (Fig. 4), there are some differences: the chemical exergy flows appear (CHj:k), the enthalpy flows (Hj:k) replace the exergy flows, and the environment (E) incorporates one small junction and one small bifurcation to indicate that it has two type of fuels and two type of products, respectively. The enthalpic (Hj:k) and the chemical (CHj:k) exergy flows of the diagram are defined based on the enthalpy (Hj and Hk) and the chemical exergy of physical flows (CHj and CHk), respectively, as shows Eqs. (11) and (12). These physical flows are defined by Eqs. (3) or (6) and (7), respectively.
H j:k = H j H k CH j:k = CH j CH k

(11) (12)

The products and the fuels of each equipment, in terms of enthalpic and chemical exergy component, are defined based on the quantity of these magnitude added to and removed from the working fluid entropy, respectively. Because the syntropy is the negative of entropy, the equipments that decrease the working fluid entropy are syntropy producers, and the others that increase the entropy of the fluid are syntropy consumers. The mathematical model is obtained by applying the Eq. (9) of cost balance to each unit of the productive diagram, as described in Section 3.1. Because the syntropy is used as an exergy component flow, the H&S Model does not use the by-product concept, i. e., each productive unit can have more than one product (the exergy component flows, separately). The flows exiting the same real unit of the productive diagram are all products, and they must have the same unit cost. Thus, the only kind of auxiliary equation is obtained by applying the common rules used (or accepted) by all thermoeconomic practitioners to formulate the auxiliary equations, i. e., the H&S Model attributes the same exergetic unit cost to all of the flows leaving the same unit. This rule is applied to the bifurcation (BHb, BPm, BSb, BPe, BHr and BSr) and also to the real units that have two exit flows: the recovery boiler (RB) and the environment (E). In this way, the H&S Model prevents the Special Consideration used by the E&S Model for cost attribution to the byproduct flows. Once that the auxiliary equations are more or less arbitrary (Tsatsaronis and Pisa, 1994) and, they are unavoidable in thermoeconomics, the H&S Model allows reducing the arbitrariness in thermoeconomics. Furthermore, it was shown recently that this Special Consideration is not applicable to any thermal cycle (Santos et al., 2008a). The H&S Model is easily applied to any thermodynamic cycle because it describes the behaviour of the system in the plane h-s. The H&S Model have been applied also to a regenerative gas turbine cogeneration plant (Santos et al., 2008b). The H&S Model considers that the residue of a combined cycle power plant has two different components: (a) the chemical component is originated in the combustion chamber, and (b) the physical (or enthalpic) component is due to the working fluid entropy increased. The thermoeconomic approaches agree that the residues must be allocated where they have been originated. Thus, the chemical component of the residue is allocated to the combustion chamber, but the enthalpic component of the residue is charged to the productive units that increase de working fluid entropy.
4. RESULTS AND DISCUSSIONS

Table 4 shows the productive internal flows, its values and its respective exergetic unit costs, considering each of the three applied models (E, E&S and H&S). In Table 4, the exergetic unit costs obtained by the E&S Model consider the Special Consideration (the byproduct concept) to attribute cost to the negentropy flow produced by the recovery boiler. All the resources entering a thermal system must be charged to the final products. The exergy of the natural gas (Qng) is the only external resource, i. e., its exergetic unit cost is 1.00 kW/kW. The combined cycle power plant has only one final product, which is the electric power. Therefore, the exergetic unit cost of the electric power flows (Pgp, Pe, Pd and Pnp) obtained by the three models (E, E&S and H&S) are the same (2.50 kW/kW). But, the exergetic unit cost of the mechanical power produced by the gas turbine (Pa /Pb) and by the steam turbine (Pc) are different for each model, because the way in which the residues and the condenser is allocated is different for the three productive structures. The exergetic unit costs of the mechanical power produced by the gas turbine (Pa /Pb) and by the steam turbine (Pc), considering the E Model, are 2.09 kW/kW and 3.19 kW/kW (respectively). The residue cost is not calculated, but it is implicitly distributed to the gas turbine and to the steam turbine (through the recovery boiler) proportionally to the value of the turbine and recovery boiler fuels (respectively). The cost involved in the operation of the condenser is allocated to the steam turbine only. If the original formulation of the Exergetic Cost Theory was considered, the residue cost would be allocated to the recovery boiler only, which would overcharge the cost of the mechanical power produced by the steam turbine (3.68 kW/kW) to the detriment of the cost of the power produced by the gas turbine (1.85 kW/kW). The E&S Model is the methodology that gets the highest exergetic unit cost of the power produced by the gas turbine (2.44 kW/kW) and consequently, the lowest exergetic unit cost of the power produced by the steam turbine (2.46 kW/kW). By using the negentropy flow joined up with the exergy flow, the gas turbine is penalized twice due to the increase of the work fluid entropy, while the recovery boiler are awarded twice due to the reduction of the work fluid
70

Proceedings of ECOS 2009 Copyright 2009 by ABCM

22 International Conference on Efficiency, Cost, Optimization Simulation and Environmental Impact of Energy Systems August 31 September 3, 2009, Foz do Iguau, Paran, Brazil

nd

entropy, because the exergy flow already contains the term (m.T0.s) that define the negentropy flow. This fact justifies why the cost of power produced by the gas turbine is overcharged to the detriment of the cost of the power produced by the steam turbine, although the E&S Model uses the Special Consideration (the by-product concept) to attribute cost to the negentropy flows. If the Special Consideration were not used, the cost of the power produced by the gas turbine would be higher (2.78 kW/kW) and the cost of the power produced by the steam turbine lower (1.75 kW/kW). There is no thermoeconomic justification to explain why the costs of the mechanical power produced by the gas turbine (in a combined cycle) is greater that the mechanical power produced by the steam turbine, once that the product of the bottoming cycle should be greater than the topping cycle products. It was previously shown (Santos et al., 2006) that, without the Special Consideration, the cost of power of a cogeneration plant obtained by the E&S Model can be so high, that contradicts the well-known thermodynamic advantage of cogeneration. Further more, the E&S Model accepts that the products of some units can be greater than their fuels and, consequently, it obtain some exergetic unit costs lower than one. In the irreversible cycles, the exergetic unit cost should be increasing along the productive structure. Table 4: Exergetic Unit Cost of the Productive Structure Flows of the Combined Cycle Power Plant PRODUCTIVE FLOW Qng E2:1 E3:2 E3:4 / E4:5 E5:1 E7:6 E8:7 E8:9 / E9:6 S2:1 / S3:2 / S4:3 S4:5 S5:1 S7:6 / S8:7 / S9:8 S9:6 H2:1 H3:2 H3:4 / H4:5 / H5:1 H7:6 H8:7 H8:9 / H9:6 CH3:2 CH5:1 Pa / Pb Pc Pgp / Pe / Pd Pnp VALUE [kW] 209,697.64 89,807.74 132,835.57 164,998.07 / 44,334.68 13,310.56 122.25 35,396.38 32,992.96 / 2,525.67 9,085.35 / 70,279.57 / 7,824.93 45,882.23 41,307.61 70.06 / 54,440.00 / 5,089.32 59,599.38 98,893.09 199,884.45 157,173.15 / 90,216.91 / 51,387.49 192.31 89,836.39 27,903.64 / 62,125.05 3,230.69 3,230.69 98,893.09 / 58,280.05 27,903.64 84,415.00 / 210.00 / 205.00 84,000.00 EXERGETIC UNIT COST [kW/kW] E Model E&S Model H&S Model 1.00 1.00 1.00 2.30 2.76 ---1.58 1.97 ---1.99 2.29 ------2.29 ---4.19 4.24 ---2.49 2.06 ------2.07 ------0.74 2.49 ---0.74 2.42 ---0.74 2.58 ---0.10 2.53 ---0.10 2.53 ------2.46 ------1.94 ------2.11 ------3.58 ------2.42 ------2.42 ------1.94 ------2.58 2.09 2.44 2.24 3.19 2.46 2.88 2.50 2.50 2.50 2.50 2.50 2.50

According to the H&S Model, the fuels of the equipment are always greater than their products and, consequently, it obtains exergetic unit costs of the internal flows that are always greater than one. The exergetic unit cost increases along the productive structure. In the analyzed plant, the exergetic unit cost of the power produced by the steam turbine (2.88 kW/kW) is coherently greater than the exergetic unit cost of the mechanical power produced by the gas turbine (2.24 kW/kW). The H&S Model takes all the advantages due to the use of negentropy in the productive structure for residue and dissipative component cost allocation, and it prevents the Special Consideration to formulate the auxiliary equations. The H&S Model productive structure shows in more details the process of residue and dissipative cost allocation by giving different and separated treatment to the chemical and enthalpic residue components.
5. CONCLUSION

This paper showed that the total exergy flow only (E Model) is not enough to define the productive structure of the combined cycle power plant representing explicitly the process of residue and dissipative component cost allocation, i.e., in the E Model the residue cost is not calculated, but it is implicitly allocated to the final products of the Brayton cycle. Furthermore, this model (E Model) does not permit the isolation of the condenser in the Rankine cycle. The introduction of the negentropy, as an internal flow of the productive structure, allows representing the residue cost distribution to the final products in details. This technique allows calculate and redistribute the cost of the residues. By using the negentropy flows, the cost of the condenser product is calculated and distributed to the productive units.
71

But, several authors apply the negentropy as a fictitious flow, joined up with the exergy flows (E&S Model), which leads to some ambiguousness because there are productive units where their fuels are less than their products. Consequently, this model obtains some exergetic unit costs less than one. To prevent excessive overcharge in the cost of the gas turbine power, the E&S Model needs a Special consideration for cost attribution to the so-called negentropy byproducts. Thus, this Special Consideration can be interpreted as a way to disguise the E&S Model ambiguousness. But this paper also showed that we can take all the known negentropy advantages by using the physical exergy disaggregated into enthalpy and syntropy, combined with the chemical exergy component (H&S Model). Thus, we prevent the ambiguousnesses regarding the fuels-products definition, and we do not need the Special Consideration. The H&S Model defines the plant productive structure according to the behavior of its subsystems in the plane h-s.
ACKNOWLEDGEMENTS

The authors would like to tank CAPES, FAPEMIG and also CNPq for the financial supports.
REFERENCES

Alves, L.G. and Nebra, S.A., 2003, Thermoeconomic Evaluation of a Basic Optimized Chemically Recuperated Gas Turbine Cycle. Int. J. Thermodynamics. Vol.6 (No.1), pp.13-22. Arena A.P. and Borchiellini R., 1999, Application of Different Productive Structures for Thermoeconomic Diagnosis of a Combined Cycle Power Plant. Int. J. Therm. Sci. (1999) 38, 601-612. Frangopoulos, C.A., 1994, Application of the Thermoeconomic Functional Approach to the CGAM Problem. Energy Vol. 19, No. 3, pp. 323-342. Frangopoulos, C.A., 1987, Thermo-Economic Functional Analysis and Optimization. Energy Vol. 12, No. 7, pp. 563571. Lazzaretto, A. and Tsatsaronis, G., 2006, SPECO: Systematic and General Methodology for Calculating Efficiencies and Costs in Thermal Systems. Energy 31 (2006) 1257-1289. Lozano, M.A. and Valero, A., 1986, Determinacin de la Exergia para Sustancias de Interes Industrial. Departamento de Termodinmica y Fisicoqumica. ETSII. Universidad de Zaragoza. Lozano, M.A. and Valero, A., 1993, Thermoeconomic Analysis of a Gas Turbine Cogeneration System. ASME Book no. H00874, WAM 1993, AES, vol. 30, p. 312-20. Santos, J.J.C.S., Nascimento, M.A.R. and Lora, E.E.S., 2006, On The Thermoeconomic Modeling for Cost Allocation in a Dual-Purpose Power and Desalination Plant. In Proceedings of ECOS 2006. Volume 1, Pages 441-448. Aghia Pelagia, Crete. Greece. Santos, J.J.C.S., Nascimento, M.A.R., Lora, E.E.S. and Martnez Reyes, A.M., 2008a, On The Negentropy Application in Thermoeconomics: a fictitious or an exergy component flow? In Proceedings of ECOS 2008. Volume 1, Pages 253-260. Cracow, Poland. Santos, J.J.C.S., Nascimento, M.A.R., Lora, E.E.S. and and Martnez Reyes, A.M, 2008b, On The Productive Structure for the Residues Cost Allocation in a Gas Turbine Cogeneration Plant. In Proceedings of ECOS 2008. Volume 2, Pages 641-648. Cracow, Poland. Torres, C. and Valero, A., 2000, Curso de Doctorado: Termoeconomia. Curso 2000-2001. Departamento de Ingeniera Mecnica. Universidad de Zaragoza. Torres, C., Serra, L., Valero, A. and Lozano, M.A., 1996, The Productive Structure and Thermoeconomic Theories of System Optimization. ME96: International Mechanical Engineering Congress & Exposition (ASME WAN 96). Torres, C., Valero, A., Rangel, V. and Zaleta, A., 2008, On The Cost Formation Process of the Residues. Energy Vol. 33. pp. 144-152. Tsatsaronis, G. and Pisa, J., 1994, Exergoeconomic Evaluation and Optimization of Energy System - Application to the CGAM Problem. Energy Vol. 19, No. 3, pp. 287-321. Valero, A., Lozano, M. A. and Serra, L., 1994, Application of the Exergetic Cost Theory to the CGAM Problem. Energy Vol. 19, No. 3, pp. 365-381. Verda, V., Serra, L. and Valero, A., 2004, The Effects of the Control System on the Thermoeconomic Diagnosis of a Power Plant. Energy Vol. 29, pp. 331- 359. von Spakovsky, M.R., 1994, Application of Engineering Functional Analysis to the Analysis and Optimization of the CGAM Problem. Energy Vol. 19, No. 3, pp. 343-364.
RESPONSIBILITY NOTICE

The authors are the only responsible for the printed material included in this paper.
72

Das könnte Ihnen auch gefallen