You are on page 1of 8

OPERATION DIAGNOSIS OF A COMBINED CYCLE BASED ON THE STRUCTURAL THEORY OF THERMOECONOMICS

Luis Correas, ngel Martnez and Antonio Valero CIRCE. Department of Mechanical Engineering University of Zaragoza Zaragoza. SPAIN Mara de Luna 3. 50015 Zaragoza. SPAIN Tel.: +34 976 76 18 63 Fax: +34 976 73 20 78 E-Mail: lcorreas@posta.unizar.es
W power (kW)

ABSTRACT Diagnosis of the performance of energy was theoretically developed based on the Structural Theory (Valero, Serra and Lozano, 1993), and traditionally Thermoeconomics have usually been applied to the design of power plants and comparison between alternatives. However, the application of thermoeconomic techniques to actual power plants has always to face the generally poor quality of measurement readings from the standard field instrumentation as an unavoidable first step. The proposed methodology focuses on measurement uncertainty estimation and performance calculation by means of data reconciliation techniques, in order to obtain the most confident plant balance upon the available instrumentation. The formulation of the Structural Theory has been applied to a combined cycle, where the Fuel-Product relationships at the component level must be optimally defined for a correct malfunction interpretation. This set of relationships determines the ability to diagnose and the level of the diagnostics obtained. The paper reports the application of the methodology to a 280 MW rated combined cycle, where performance diagnosis is illustrated with results from a collection of actual operation data sets. The results show that data reconciliation yields sufficient accuracy to conduct a thermoeconomic analysis, and how the estimated impact on fuel correlates with physical causes. Hence the feasibility of thermoeconomic analysis of plant operation is demonstrated. NOMENCLATURE Scalars (upper case) B extensive exergy (kW) F exergy fuel of a thermal system (kW) I irreversibility (kW) L losses (kW) LHV lower heating value (kJ/kg) MF malfunction (kW) P exergy product of a thermal system (kW) Q heat (kW) 1

Scalars (lower case) b intensive exergy (kJ/kg) h enthalpy (kJ/kg) k unit exergy cost (tpu) m mass flow (kg/s) n number of subsystems uncertainty s entropy (kJ/kg/K) u vector of measured variables v vector of unmeasured variables w measurement weighting x solution vector Greek letters efficiency (tpu) specific consumption of each fuel in a thermal system Lagrange multipliers Subscripts b exergetic 0 referred to environment s negentropy T referred to total Superscripts * Exergy cost 0 reference state of calculus

INTRODUCTION Performance monitoring and diagnosis of thermal systems has become a very important topic in recent years, revealed by the many existing plant monitoring packages commercially available. These software applications try to achieve various goals, like instrument fault monitoring, energy and cost accounting, or production and maintenance scheduling among others. In addition to dispense information at the corresponding levels, a desirable objective is to detect deviations from the production objetives. In this sense, thermoeconomic analysis, a second law based method, correctly allocates and evaluates efficiency losses (irreversibility increase). It is the aim of this paper to demonstrate the feasibility of applying thermoeconomic analysis as a tool for performance enhancement in power plants. Although thermoeconomic diagnosis had already been theoretically developed (Lozano, Bartolom, Valero, Reini, 1994), the main problem when trying to diagnose the operation of an existing plant is that the available information is limited, in quantity as well as in accuracy, so that the uncertainty propagates to the final results, and must be accounted for. It is important to remark that this study is based on an existing combined cycle, IGCC Puertollano, and nearly 400 actual data sets have been collected and pretreated to tune up the performance calculation modules included in a monitoring software (Valero, Correas, and Serra, 1998). The paper is structured into three main topics, the first one devoted to performance calculations from measurement readings. A first analysis of the field instruments detects the weaknesses and forecasts the expected uncertainty. Through the study of historic data, the first uncertainty evaluation is tuned using data reconciliation techniques, and hence the measurements can be weighted accordingly. This procedure yields the best possible plant balance upon the available instrumentation. Although there are not many redundant measurements in power facilities, the positive effect of this calculation procedure is clearly pointed out. How the Structural Theory (Valero, Serra, and Lozano, 1993) has been applied to the combined cycle is the next item. The plant is represented as the so called productive structure, that fulfills some syntactic rules, and that helps to relate component malfunctions or operation strategies with the local increase in fuel consumption. In our case, a simplified productive structure of the combined cycle is explained, focusing on those components that are more likely to degrade. The definition of fuels and products is of utmost importance to clearly locate the malfunctions and its increase on fuel consumption. Once the measurement uncertainties have been assessed and given the productive structure that best serves our diagnostic purposes, a sensitivity analysis is conducted at the level of single impact on fuel. Finally the paper illustrates the true capabilities of operation diagnosis from actual data sets. OVERVIEW OF THE STRUCTURAL THEORY Thermal power plants can be depicted as a set of components, which are joined by material and energy streams. These components interact with their environment, consuming some external 2

resources, which are then transformed into certain products. The final purpose of this transformation is to increase the economic usefulness. Exergy permits proper quantification of the thermal quality of the energy, so exergetic analysis must be the concept base line to begin any study about installations in process engineering (Lozano and Valero, 1988). It is possible to translate the information of a plant defined by a flow sheet of physical streams, into an exergy conversion processes which consumes exergy resources (fuels of the system) and produce others transformed by the process itself (products of the system) (Torres, Valero and Serra, 1999). This information may be expressed in a fuel-product graph. composed by technical units that represent the processes in the plant (heat and mass transfer equipment, turbines, compressors, pumps, pipes and physical junctions), exergy mixers or structural junctions, in which exergy produced in different parts of the plant are joined, and structural branches, in which exergy is split to be consumed in several components. It is important to remark that in exergy mixers and in branches there is no exergy destruction and these ones do not represent physical components in the plant, but they serve to distribute exergy between the different technical units comprising the plant (Lozano, Bartolom, Valero and Reini, 1994). For instance, in Figure 1, is represented a system which is composed of four technical units represented by squares, one junction represented by a rhombus, one branch represented by a circle, and one physical junction represented by a special symbol E3, where the streams P13 and P23 are mixed. The product of a physical junction represents an actual mixing process of two or more physical streams as it is, for example, mixing between the low pressure steam and intermediate pressure steam in the cross-over pipe in a steam turbine. Instead of this, the product of a structural junction represents the sum between two or more structural streams in the productive structure, for example in Figure1, the exergy content in stream P70 is the sum of the exergy content in stream P57 and in stream P67. The irreversibility of a component i is:

I i = Fi Pi = Pji Pi
j =0

(1)

where Pji is the exergy produced by the component j to be consumed in the component i to produce Pi . Note that by measuring the irreversibility, all the possible internal losses in the component are taking into account. Therefore, any perturbation, internal malfunction or deterioration in a component can subsequently be measure as:

I i = I i I i0 = Component Perturbation

(2)

because locally, there is no loss neglected. This is probably the strongest reason for using exergy as a tool for diagnostic purposes. When analyzing inefficiencies of a system a reference state () must be established, and the deviations of the performance from

this state can be calculated. Thus, in practice, the exergy savings can be expressed like increments of inefficiency between operation and design. If we define the specific fuel consumption ji for each fuel j feeding the component i as ji = Pji / Pi , it is possible to relate it with internal malfunctions through:
n n

MFi = Pi 0 ji = MF ji
j =0 j =0

(3)

Now, the relationship between that local perturbation and the overall impact on additional exergy resources FT ,i , needed to compensate that malfunction can be obtained by introducing the concept of exergetic cost. The unit exergetic cost k * j of a flow j in a given plant is the amount of the exergy resources needed to produce it (Valero, Lozano, Muoz, 1986). Therefore, if the total production of the plant is constant and there are no other induced malfunctions caused by the component s deterioration then:

ited by the gas dew point to 60 C, in order to prevent corrosion. This is achieved by mixing feed water from the deaerator. A preliminary instrumentation analysis based on acceptance test procedures gives an overview of the quantity and quality of field instruments. The expected uncertainty of every measurement, assuming a standard calibration has already been made, is composed from the various error sources, that is, device (the instrument itself), measurement method (influence of corrections or second order effects) a and signal transmission chain, where estimations are taken from literature and technical specifications (Deieck, 1992; Spink, 1972; Benedict, 1984).
2 2 2 measurement = method + device + transmissi on

(6)

Gas turbine instrumentation has been satisfactorily compared with norms ISO 2314 (1989) and ASME PTC 22 (1985) (see Table 1 for details). As the norms state, a direct measurement of the turbine inlet temperature (TIT) and air mass flow are not feasible. These can be accurately estimated by calculating the energy balance around the gas turbine.
Measurement number of readings Compressor inlet temperature Compressor inlet pressure Compressor outlet temperature Compressor outlet pressure Natural gas mass flow 4 1 2 1 1 1 1 avg. of 6 1 3 C 0.5 % 3 C 0.5 % 0.5 % 2.3 C 0.15 bar Avg. 1 % 0.3 % expected uncertainty

FT ,i = k MF ji MFi
j =0 * j

(4)

If a local malfunction also provokes induced malfunctions in the other components, we must take into account that and the final formula becomes:

FT = FT ,i =
i =1

i =1 j = 0

k P
* j

ji i

(5)

Natural gas temperature Natural gas pressure Average TOT Power output

as it is defined in Torres et al. (1999). Consequently, once the individual contributions on fuel impact each component of the plant have been calculated, it is possible to diagnose which components offers potential for efficiency improvement. PERFORMANCE TEST OF THE COMBINED CYCLE The combined cycle in which this paper focuses consists of a single gas turbine and one steam turbine, with a three-pressure level heat recovery steam generator. Nominal output exceeds 280 MW when firing natural gas, where the gas turbine produces 70 % of the total power. The gas turbine is a Siemens V94.3, with dual fuel burners. It achieves nearly 37 % efficiency, with a firing temperature of 1120 C, while yielding 560 kg/s of hot gases to the HRSG at approximately 540 C. Compressor inlet guide vanes (IGVs) regulation allows to keep turbine outlet temperature high at part load so as not to degrade steam cycle performance. The three pressure levels of the HRSG consist of economizers (except the low pressure one), evaporators and superheaters. High pressure superheater and reheater show interleaved arrangement. Live steam conditions are 80 bar and 520 C. Exhausted gases exit the HRSG at about 110 C, just after that the low temperature sensible heat has been used for condensate preheating. The lower end temperature of this preheater is lim-

Table 1.- Gas turbine instrumentation

mair hair _ in + m fuel LHV + mDeNOx hDeNOx + W aux = Woutput + m gas hgas _ out + Qcooling + LGT
(7)

A further energy balance around the combustion chamber yields a good estimation of the TIT.

mair hcompressed _ air + m fuel LHV + mDeNOx hDeNOx = m gas hgas (TIT ) + Lcc
(8)

The most critical measurement are those related with the fuel. A turbine flow meter with an accuracy of 0,5 % constitutes the main error source of the whole gas turbine, while the natural gas composition does not represent a great problem because of its stability over time. In addition, two daily analysis, one from the supplier and one from the laboratory provide confident data. Turbine outlet temperature (TOT) is probably the most complicated measurement, and every manufacturer applies different technical solutions. In this case, the sought thermodynamic temperature is the average of an array of six double thermocouples, arranged in a non symmetrical fashion in the exhaust gas 3

duct. The sensors are located as near as possible to the exhaust flange, so as to shorten reaction times for the sake of reliable and safe operation, although individual readings in different gas paths could deviate up to 30 C from the average. Once the air mass flow can be estimated from energy balance, all other credits (DeNOx steam) and debits (heat extracted from turbine blade cooling air) can be accurately calculated from field measurements. The steam cycle has been analyzed according to norms ASME PTC 4.4 (1981) and ASME PTC 6 (1976). Mass flows of individual feed water streams and make-up water are measured with orifices. Additionally, live steam mass flow is directly measured, as well as DeNOx steam, the two low pressure steam mass flows (to the deaerator and to the cross-over pipe), main condensate and condensate recirculation to the preheater, all by means of orifice plate, corrected in pressure and temperature. This permits some redundancy that could serve whether to check results or to reconcile data. Some minor streams must be estimated from mass balances, as they are the high pressure steam to DeNOx, or the preheater bypass. Also, reheat mass flow is obtained by deducing an estimation of seal steam. Pressure and temperature upstream and downstream of every heat exchanger and steam turbine section are provided, so that water-side energy balances can be closed. Four manometers and one absolute pressure measurement are installed in the condenser. A simplified list of instrumentation is as follows in Table 2:
Measurement Mass flow measurements Live steam temperature Live steam pressure Reheat steam temperature Reheat steam pressure Condenser pressure Power output number of readings 9 4 9 4 5 1 expected uncertainty 3-5% 5.5 C 0.4 1.2 bar 5.5 C 0.15 0.6 bar 0.8 % 0.3 %

Table 2.- Steam cycle instrumentation

The temperature profile inside the HRSG can only be reconstructed from the water side, because from the gas side there are only four thermocouples at the stack (4 C individual uncertainty). It is obvious the technical challenge of accurately measuring gas temperature inside the HRSG, that otherwise it is not of prime importance.
Result Gas turbine Efficiency Compressor isentropic efficiency Expander isentropic efficiency Air mass flow Turbine inlet temperature uncertainty 1 0.2 % 0.6 % 0.3 % 6.3 kg/s (1.1 %) 5.7 K (0.5 %)

Once it has been assessed that the field instrumentation provides sufficient and accurate enough readings, a performance test code is developed in order to calculate not only the main parameters of contractual relevance, such as efficiencies, but also every stream in gas and water cycles. Only if every temperature, pressure and mass flow were calculated, it would be possible to observe how the single components perform. This is the concept of performance test used in this methodology. As it has been already mentioned, the instrumentation in the gas turbine suffices to close the balances, but does not offer any redundancy to check with. Air mass flow and turbine inlet temperature are calculated with expressions (6) and (7). The results of the performance calculations are highly sensitive specially to the fuel input, and the turbine outlet temperature (see Table 3). In the other hand, the steam cycle has a certain amount of redundant measurements (11 independent ones), that allow application of an uncertainty minimization algorithm instead of a sequential procedure. The sequential approach uses the strictly minimum set of data, and closes mass and energy balances stepwise in every component, solving an unknown in each step. Once the calculation is finished, the rest of the instrument readings serves to check up how close is the solution to the actual data. After a correction of the input data, a new calculation is then performed, until the check points lay under a given tolerance. Nevertheless, it is difficult to formalize the stop criterion and how the actual data should be corrected, which is normally done by rules of thumb and one by one. There are also other physical constraints for the problem, like temperature differences at the heat exchangers and the isentropic expansion, specially in the low pressure section. A sequential procedure would transmit all the measurement errors to the non instrumented points, like the condensing steam, whose quality is never measured. Normally, one can not rely too much on a few, very good measurements, but on many of reasonable accuracy, some of them being redundant. Because the measurements are subject to random and systematic error, they are not normally consistent with the redundancy within themselves. The set of instrument readings are reconciled by minimizing a weighted sum of squares of the normalized differences between the measured and calculated values while satisfying the modeling constraints. All equipment and stream variables are stored in the vector x of unknowns, of length M:

x = (u: v ) T

(9)

where u and v are the subsets of the measured and unmeasured variables, having lengths P and M-P respectively. The problem can then be formulated as: minimize

g (u) = wi (ui ,meas ui ,calc ) 2


i =1

(10)

Table 3.- Gas turbine performance test results


1

subject to a set of N nonlinear independent constraints (mass and energy balances, and component specific equations):

Uncertainty is expressed in the same units as the result, i.e. the efficiency of the gas turbine is 37.0 0.2 %. 4

f ( x) = 0

(11)

The weighting factor wi is an arbitrary value of the relative confidence of the measurement, assumed as the estimated uncertainty. The optimum set for the measurement readings is found by solving the Lagrangian (For more details on the implementation, see Stephenson and Shewchuck, 1986, and Heyen, Marechal and Kalitventzeff, 1996).
T G ( x, ) = g (u ) + f ( x)

(12)

Table 4 presents the uncertainty on some parameters, calculated with a sequential procedure and compared to the uncertainty obtained with the data reconciliation algorithm. Global parameters like combined cycle efficiency, that depend on few measurements, can not be enhanced.
Result uncertainty sequential pr. Combined cycle Efficiency HP isentropic efficiency IP isentropic efficiency LP isentropic efficiency HRSG efficiency Total heat transferred in HRSG Heat transferred in Reheater 0,25 2,8 2,7 9,3 3,2 11 1,5 d. reconciliation 0,25 1,0 1,1 3,3 0,8 3 0,5 % % % % % MW MW Units

Table 4.- Comparison of results from sequential procedure and data reconciliation.

STRUCTURAL THEORY APPLIED TO COMBINED CYCLES The physical structure of combined cycle can be represented as a graph of exergy consumptions and productions in the so called productive structure. There are innumerable graphs for a given physical structure, so many as different fuel and product definitions for every component. The usual cost assignment rules, like exergetic cost theory, average costing, or the last-infirst-out method (Lazaretto and Tsatsaronis, 1997) can be formulated by means of the productive structure (Erlach, 1998). The simplified productive structure depicted in Figure 2 is proposed, where only the most determinant exergy streams have been taken into account, hence ignoring seal steam and blade
Component Compressor Expander Combustion Chamber Fuel

level, because it is a second order source of performance degradation. Nevertheless, formal cost assignment has been carried out, where fuel for any heat exchanger is the exergy decrease in the gas side, and the product, the exergy increase in the steamwater side. In few words, the productive function of the turbines is to generate shaft work from an exergy decrease, while the nonexhausted outlet flow is later used in another component. The conversion of work into exergy in the compressor is considered as another fuel of the expander. The combustion chamber raises the temperature, and hence the exergy of the air, but has little influence on pressure losses, that have been neglected. The product of the combustion chamber is diverted proportionally to the expander and to the HRSG. The HRSG produces various steam streams from condensate, each with a different content on specific exergy, affected by different heat transfer processes, and hence with a different exergetic cost. Finally, the condenser is responsible for closing the watersteam cycle, however condensate is irrelevant from a purely exergetic point of view (null exergy, not represented). Entropy decrease, that is, negentropy generation represents conveniently the various effects concerning condenser performance: From a cost formation point of view, the low exergy steam streaming out from the low pressure section is a minor exergetic loss in the cycle (about 50 kJ/kg, 6 % of the total exergy losses in the steam cycle), being condensate at ambient temperature the dead state. This exergy loss must be accumulated to the cost of the final product of the steam cycle, where the usual criterion is to add negentropy proportionally to entropy generation in every component. This assumption does not affect the cost of the final product. Furthermore, it permits accounting for capital costs of the condenser if necessary. When detecting malfunctions, a higher temperature difference of the condensate versus the ambient is only explained by a worse heat exchange in the condenser. The greater exergy of the exhausted steam, evaluated to the cost of low pressure steam, the greater is the impact on fuel consumption:
Product

Wcompressor
mGas (bGasInGT bGasOutGT )

m Air bAirOutCompressor Wcompressor + WOutput mGas bGasInGT m Air b AirOurCompressor

Steam turbine Condenser

mSteam (bSteamIn bSteamOut )


m ExhSteam bExhSteam

m NaturalGas b NaturalGas mDeNOx bDeNOx

WSteamTurbine
m ExhSteam T0 (scondIn scondOut )

Table 5.- Definitions of Fuels and Products.

cooling air. Moreover, the HRSG appears at a high aggregation 5

* * MFcondenser = k LPsteam condenser Pcondenser * k LPsteam BexhSteam

uum pressure, rises with respect to ambient (or cooling water) temperature. If the work done by the LP turbine is:

(13)

WLPST = msteam (hin hout )

(16)

condenser

B B = exhSteam exhSteam Bs real Bs

(14)
ref

then, applying the definition of isentropic efficiency and supposing constant inlet conditions (and hence negligible effects on s):
0 WLPST = msteam s (hout ,s (Tcond ) hout ,s (Tcond ))

Pcondenser = Bs = mexhSteam To (sin sout )


Fuel/Product Exergy of hot gases Exergy of compressed air Gas turbine power HP steam IP steam LP steam Reheat DeNOx steam Steam to IP turbine Steam to LP turbine and condenser HP turbine power IP turbine power LP turbine power Negentropy cost 1,57 2,02 1,92 1,93 1,89 2,31 1,77 1,94 1,90 2,00 2,05 2,00 2,30 0,07

(17)

(15)

or,

W LPST = m steam s

hout , s Tcond
Tin , Pin

dTcond

(18)

Table 6.- Unit exergy cost

If heat transfer decreases and hence condenser temperature and pressure raises, the low pressure turbine will produce less power, that should be accounted for as a higher fuel consumption. The cost in terms of additional natural gas can be approximated as the decrease in power multiplied by its unit cost (approximately 2.3). Figure 3 shows that the relationship obtained between impact on fuel in the condenser and the estimation of loss of work in the LP turbine matches the cost of the power (with a 5% allowance). This result clearly indicates that the procedures for cost assignment as well as for impact detection work properly, even at the end of the productive chain, as an empirical validation. The influence of other parameters affecting, for instance the compressor or the turbines, has been detected and could be explained in a similar way,(see Figures 5 and 6).
Component Compressor Compressor outlet temperature HP Steam turbine HP turbine outlet temperature Live steam pressure 1 bar 1 C 24 kW 17 kW 1 C 33 kW 1 C 301 kW Measurement Measurement error Apparent malfunction

The only assumption made when designing this productive structure was that a component using or processing but not exhausting a flow (i.e. turbines), did not affect the exergetic cost of the outgoing stream. That is, in our case, the whole irreversibility was added to the shaft power. As every component has only one product, no further assumptions on costs are then necessary. Table 5 expresses the fuel-product model applied, and Table 6 the unit exergy cost obtained. DIAGNOSIS OF MALFUNCTIONS Individual component performance can be expressed as a function of some internal parameters, which allow degradation monitoring. The productive structure introduces the information of how every malfunction affects overall efficiency by means of the exergetic cost. Translating into economic cost, it becomes a valuable indicator to decision making. Figures 3, 4, 5 and 6 have been generated from the systematic comparison of nearly 400 data sets, corresponding to natural gas operation in the whole load range. One of the sets has been selected as best feasible operation, and the rest of them compared to it. Some effects are presented, that relate physical parameters with the increased consumption of natural gas. The increased fuel consumption due to a malfunction of the condenser can be easily interpreted as a loss of work done by the low pressure section when condensate temperature, that is vac-

Live steam temperature Condenser Condenser pressure

1%

1138 kW

Table 7.- Sensitivity analysis on impact on fuel.

The above mentioned graphics show a good repeatability of malfunctions because of the relative low scatter. Nevertheless, the results of impact on fuel are very sensitive to potential errors from measurements, as it can be seen in Table 7. The global impact on each analyzed component is determined only by a few measurements. If a certain error affects the measured value, a fictitious impact on fuel will be calculated, that could mislead to an incorrect interpretation and subsequent corrective actions. Although mathematically right, the results must always be criticized from the point of view of instruments quality.

CONCLUSIONS It has been demonstrated that data reconciliation helps considerably to reduce uncertainty of component performance, whenever a certain amount of redundant instrumentation is available, although the precision of global efficiency can not be enhanced. The simplified productive structure allows a good allocation of malfunctions and a reasonable estimation of the increase on fuel consumption. A more complicated structure appears as unnecessary within the scope of the paper for demonstrative purposes, although a detailed application to an actual case does need to take into account the complete layout of the plant to achieve a coherent exergy and cost balance. The results obtained from many actual data sets are encouraging for further developments. Repeatability is fairly good, although sensitivity of the results on the single instruments indicates that the main problem when applying the Structural Theory for monitoring purposes is measurement quality. This study has been conducted in the frame of the development of a monitoring system for the operation diagnosis based on thermoeconomic techniques, that has been reported by Valero et al., (1998). The present paper shows the feasibility of applying Thermoeconomics to actual systems, with reliable and useful results. The development will continue further in two trends. First, the monitoring system is being expanded to cover the whole IGCC plant. In the other hand, the aim of the development is to yield conclusions or suggestions to the personnel of the various departments (operation, production, maintenance), that are not familiar with exergetic analysis, in a fashion that could serve their daily activities. The true potential of thermoeconomic analysis must now be exploited, once the credibility of the results has been demonstrated.
Exergy of hot gases

FIGURES
P51

P01

E1
P13

P45

E5
P57

E3
P34

E4

E7

P70

P23 P02

P67 P46

E2

E6

Fig 1.- Example of Productive Structure

25000

20000

y = 2,36x 2 R = 0,99
15000

10000

5000

-5000 -2000 0 2000 4000 6000 8000 10000 12000

Fig 3.- Impact on fuel [kW] in the condenser related with the loss of work in the low pressure steam turbine [kW].

Negentropy HP steam DeNOx Steam outlet steam Negentropy Exergy consumption HPST Shaft power

Heat Recovery Steam Generator

Natural Gas

Combustion Chamber

IP steam

Exergy consumption

IPST

Shaft power

reheat

outlet steam Negentropy Exergy LPST Shaft power

Exergy of compressed air

LP steam

Exergy consumption

Exhausted steam

Power

Power Negentropy

Output Power

Negentropy

Fig 2.- Productive Structure of the Combined Cycle

12000

10000

8000

6000

4000

2000

-2000

-4000 -4 -3 -2 -1 0 1 2

Fig 4.- Impact on fuel [kW] in the condenser due to temperature difference between condensate and ambient temperature [K]
1400 1200

1000

800

600

400

200

-200 -4,0 -3,5 -3,0 -2,5 -2,0 -1,5 -1,0 -0,5 0,0 0,5

Fig 5.- Impact on fuel [kW] due to variation on isentropic efficiency of the high pressure steam turbine [%]
12000 10000

8000

6000

4000

2000

-2000

-4000 -4 -3 -2 -1 0 1 2

Fig 6.- Impact on fuel [kW] due to variation on compressor isentropic efficiency [%]

ACKNOWLEDGMENTS The authors would like to acknowledge the management of Elcogas power plant for their faithful collaboration at every moment, specially to Dr. Manuel Trevio, chief executive officer of the company, and Dr. Javier Pisa, head of the Technology Group and Ing. Ana Glvez, for their constant and active support to this idea. Last but not least, we acknowledge all the coworkers and researchers within the development of the thermoeconomic diagnosis, Ing. Pedro Casero, Ing. Luis M. Blasco, Ing. Jos L. Milln and Dr. Scarpellini and the collaboration of Dr. Luis Serra and Dr. Csar Torres. REFERENCES ASME, 1976, Performance Test Codes: Steam Turbines, PTC 6. The American Society of Mechanical Engineers. 8

ASME, 1981, Performance Test Codes: Gas Turbine Heat Recovery Steam Generators, PTC 4.4. American Society of Mechanical Engineers. ASME, 1985, Performance Test Codes: Gas Turbine Power Plants, PTC 22. American Society of Mechanical Engineers. Benedict, R. P., 1984, Fundamentals of Temperature, Pressure and Flow Measurements, 3rd Ed. John Wiley and Sons. British Standard, 1989, British Standard Specification for Gas Turbine acceptance tests. ISO 2314. Deieck, R., 1992, Measurement Uncertainty. Instrument Society of America. Erlach, B., 1998, Comparison on Thermoeconomic Methodologies: Structural Theory, AVCO and LIFO, Application to a combined cycle. ECOS 98. Heyen, G.; Marechal, E.; Kalitventzeff, B., 1996, Sensitivity Calculations and Variance Analysis in Plant Measurement Reconciliation, Computers Chemical Engineering Vol. 20 Sup. S539-S544. Lazaretto, A.; Tsatsaronis, G., 1997, On the Quest for Objective Equations in Exergy Costing, AES-Vol. 37, Proceedings of the ASME. Advanced Energy Systems Division, pp. 197-210. Lozano, M. A.; Bartolom, J.L.; Valero A.; Reini, M., 1994, Thermoeconomic diagnosis of energy systems, Flowers 94, Florence World Energy Research Symposium. July 6-8, Florence, Italy. Lozano, M. A.; Valero, A., 1988, Methodology for Calculating Exergy in Chemical Processes in Thermodynamic Analysis of Chemically Reactive Systems, pp. 77-86, W. J. Wepfer, G. Tsatsaronis and R. A. Bajura eds., The American Society of Mechanical Engineers, New York, NY. Spink, L. K., 1972, Principles and Practice of Flowmeter Engineering, 9th Ed. The Foxboro Company. Stephenson, G.R.; Shewchuk, C.F., 1986, Reconciliation of Process Data with Process Simulation, AIChE Journal, Vol. 32, No. 2, pp. 247-254. Torres, C.; Valero, A.; Serra, L., 1999, Structural Theory and Thermoeconomic Diagnosis. Part I: On Malfunctions and Dysfunction Analysis, ECOS 99. Valero, A.; Correas, L.; L. Serra, 1998, On-Line Thermoeconomic Diagnosis Of Thermal Power Plants. NATO ASI on Thermodynamics and the Optimization of complex Energy Systems, July 1998, Constantza, Rumania. Valero, A.; Lozano, M. A.; Muoz M., 1986, A General Theory of Exergy Saving: PartI, On Exergetic Cost, PartII, On the Thermoeconomic Cost, PartIII, Exergy Saving and Thermoeconomics in Computer-Aided Engineering of Energy Systems, Vol. 3-Second Law Analysis and Modelling, pp. 1-22, R.A. Gaggioli ed.,.The American Society of Mechanical Engineers, New York, NY Valero, A.; Serra, L.; Lozano, M. A., 1993, Structural Theory of Thermoeconomics, ASME Winter Annual Meeting Symposium on thermodynamics and the design, analysis and improvement of energy systems. November 28-December 3, Session II General Thermodynamics & Energy Systems.

You might also like