26.07.2013 Views

Thermoeconomic analysis and simulation of a combined... - circe ...

Thermoeconomic analysis and simulation of a combined... - circe ...

Thermoeconomic analysis and simulation of a combined... - circe ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

THERMOECONOMIC ANALYSIS AND<br />

SIMULATION OF A COMBINED POWER<br />

AND DESALINATION PLANT<br />

Departamento de Ingeniería Mecánica<br />

Ph. D. Thesis<br />

Francisco Javier Uche Marcuello<br />

Universidad de Zaragoza


THERMOECONOMIC ANALYSIS AND<br />

SIMULATION OF A COMBINED POWER<br />

AND DESALINATION PLANT<br />

Departamento de Ingeniería Mecánica<br />

Universidad de Zaragoza<br />

Ph. D. Tesis<br />

Francisco Javier Uche Marcuello<br />

Zaragoza, Mars 2000


D. Antonio Valero Capilla, Catedrático del Departamento de Ingeniería Mecánica<br />

de la Universidad de Zaragoza, y D. Luis Serra De Renobales, Pr<strong>of</strong>esor Titular del<br />

Área de Máquinas y Motores Térmicos de la Universidad de Zaragoza<br />

CERTIFICAN<br />

que la memoria titulada <strong>Thermoeconomic</strong> Analysis <strong>and</strong> Simulation <strong>of</strong> a Combined<br />

Power <strong>and</strong> Desalination Plant presentada por el Ingeniero Industrial<br />

D. Francisco Javier Uche Marcuello para optar al grado de Doctor en el programa de<br />

Optimización Energética del Departamento de Ingeniería Mecánica, ha sido realizada<br />

bajo su dirección.<br />

Zaragoza, 20 de Marzo de 2000<br />

Fdo: Antonio Valero Capilla Fdo: Luis Serra de Renobales


a Sonia


Agradecimientos<br />

Quiero agradecer especialmente la realización de esta tesis doctoral a mis padres Luis<br />

y Pilar, y a mi hermano José Luis por su paciencia y ánimos para realizarla, a pesar de<br />

no entender a veces muy claramente la finalidad de la misma.<br />

Por supuesto, Natalia es la que más me ha tenido que aguantar y animar en los malos<br />

momentos que a veces he tenido. Además, ella ha tenido siempre un interés especial<br />

para que yo la realizara.<br />

Los directores de mi tesis, Antonio y Luis, han estado siempre a mi lado disponibles<br />

para cualquier duda o sugerencia en su realización. Nuestras reuniones periódicas han<br />

servido para enriquecerme personalmente. Esta tesis también ha servido para establecer<br />

una relación especial de amistad y confianza con Luis, que para mí es fundamental<br />

en el trabajo diario.<br />

También quiero agradecer al personal de la Central Térmica Teruel (ENDESA) por su<br />

flexibilidad de horarios, que me ha permitido desarrollar gran parte de mi tesis doctoral<br />

durante mi estancia en Andorra. Y a mis compañeros de piso durante dicha estancia,<br />

que me dejaron trabajar en todo momento sin impedimento alguno.<br />

Finalmente, quiero agradecer a Rosa y a Morris su ayuda en la edición. Y a esa gran<br />

familia que es CIRCE, y al gran ambiente que existe dentro de ella.<br />

Acknowledgements<br />

The financial support provided by ICWES (International Center for Water <strong>and</strong> Energy<br />

Systems, United Arab Emirates) is gratefully acknowledged. Sincere appreciation is<br />

expressed to D. M. K. Al-Gobaisi, Director <strong>of</strong> ICWES, for his continued support <strong>and</strong><br />

encouragement during the course <strong>of</strong> this thesis. The discussions that the author <strong>and</strong><br />

my directors had with him <strong>and</strong> Ali El-Nashar <strong>and</strong> Asghar Husain were very helpful.<br />

Thanks are also extended to Hanif Sultan <strong>and</strong> John Nynam who provided the technical<br />

information essential to the design <strong>of</strong> my simulator.


Resumen<br />

La desalación de aguas de mar o salobres es una de las formas más utilizadas para<br />

dotar con la calidad suficiente a la población de los recursos hídricos necesarios para<br />

su manutención y desarrollo. En un sector industrial en constante crecimiento, ya que<br />

el consumo humano per cápita sigue aument<strong>and</strong>o constantemente con el incremento<br />

del nivel de vida, a pesar de las campañas busc<strong>and</strong>o el ahorro y la racionalidad en el<br />

consumo, sobre todo en la agricultura intensiva.<br />

España es país que cuenta con un claro déficit de agua en las zonas costeras del<br />

Levante y Sur, así como en los dos archipiélagos principales (Baleares y Canarias),<br />

dichas zonas coinciden con ser las más turísticas del país, lo que significa que la<br />

dem<strong>and</strong>a se multiplica en verano. Sin tener en cuenta la posibilidad de efectuar trasvases<br />

de otras cuencas hidrográficas no deficitarias, el problema está siendo resuelto<br />

principalmente por plantas de Osmosis Inversa, plantas cuyas dimensiones y producción<br />

se adecuan mucho mejor a las necesidades de los diferentes tamaños de los<br />

núcleos ó asentamientos estables de población. El coste del agua producida sigue<br />

siendo muy alto en comparación con la obtención por medios naturales, pero sin<br />

embargo es menor que otros métodos de desalación.<br />

Sin embargo, la situación de España no es extrapolable a las zonas con verdaderos<br />

problemas de escasez de agua: los países desérticos del Golfo Pérsico. Su escasísima<br />

pluviometría, sus elevadas temperaturas durante todo el año y la casi nula impermeabilidad<br />

de sus suelos disparan su consumo de agua. Son además países de relativamente<br />

reciente creación, por lo que la dem<strong>and</strong>a de energía eléctrica también debe<br />

ser resuelta. La instalación de gr<strong>and</strong>es plantas de cogeneración permite a la vez resolver<br />

los dos problemas, con la utilización de los inmensos recursos petrolíferos y gas<br />

de la zona. Las plantas duales de generación de potencia acopladas con las unidades<br />

de desalación por destilación flash multietapa producen el 80% del agua desalada en<br />

el mundo. Pero ello no significa que sea el método más eficiente de producir esos dos<br />

productos necesarios para toda sociedad.<br />

El análisis termoeconómico permite conocer el funcionamiento interno de dichas<br />

plantas de generación de electricidad y agua dulce, las posibilidades de ahorro que<br />

<strong>of</strong>rece este modo combinado de producción. Es esencial realizar dicho análisis de<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


12<br />

Resumen<br />

forma conjunta, cosa que normalmente no se hace en este tipo de instalaciones: cada<br />

planta es gestionada independientemente.<br />

Esta Tesis Doctoral desarrolla el análisis termoeconómico completo de la planta de<br />

cogeneración más gr<strong>and</strong>e que actualmente existe (en cuanto a la producción de agua<br />

por unidad desaladora), que consta de una planta con una turbina de vapor para la<br />

generación de electricidad y una desaladora por destilación flash de un único efecto<br />

por cada una de sus etapas. Es una tesis eminentemente práctica, es decir, trata de<br />

aplicar las metodologías que la Termoeconomía actualmente está aplic<strong>and</strong>o a otros<br />

sistemas tales como plantas de potencia a un sistema muy complejo en el cual los<br />

procesos químicos también son importantes en el balance de la instalación, no sólo<br />

los procesos mecánicos y térmicos.<br />

El análisis termoeconómico comprende cuatro partes principales que se detallan a<br />

continuación:<br />

• En primer lugar, el análisis de costes permite conocer los costes físicos de los flujos<br />

más importantes de las dos plantas, así como los costes finales de producción<br />

de agua y energía, teniendo en cuenta los costes de operación y de adquisición y<br />

mantenimiento de los equipos de la planta. Dicho análisis se basa en la creación<br />

de un modelo termoeconómico que representa de una forma funcional los procesos<br />

que ocurren dentro de la planta de potencia y de agua. Los resultados obtenidos<br />

son comparados con métodos tradicionales de contabilidad de costes que se<br />

han usado para asignar costes a los productos industriales.<br />

• Después, el análisis desarrolla el diagnóstico de la planta combinada, es decir,<br />

analiza los efectos provocados por una o varias ineficiencias simuladas dentro de<br />

la planta. Para ello, se ha construido un simulador de los dos procesos a partir de<br />

un modelo matemático y datos reales de una planta de cogeneración, que permite<br />

conocer los estados termodinámicos de referencia y con la ineficiencia con una<br />

precisión suficiente para nuestro análisis. Dichos efectos se traducen a un consumo<br />

adicional de fuel, incremento en la irreversibilidad de los diferentes procesos<br />

y una menor eficiencia en los mismos, además de ayudar a conocer las relaciones<br />

de los diferentes componentes de una instalación. En este análisis se demuestra<br />

que la planta de potencia los parámetros guía de funcionamiento de cada componente<br />

son locales, es decir, una variación de ellos no significa prácticamente al<br />

resto de componentes del sistema. Sin embargo, en la unidad MSF todos elementos<br />

principales están interconectados a través de los flujos principales que circulan<br />

por los destiladores, y por lo tanto los fallos ó mejoras sufridas en el<br />

funcionamiento de la planta afectan a toda ella, no sólo al equipo en el que están<br />

ocurriendo.<br />

• La tercera parte del análisis termodinámico es la optimización de la planta de potencia<br />

a partir de la optimización local de sus componentes. En la planta destiladora<br />

de agua la optimización local no es posible al no estar sus componentes<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Resumen<br />

termodinámicamente aislados, como ya se vió en la diagnosis de la planta. Esta<br />

metodología es muy valida para el diseño de nuevas plantas o la readaptación de<br />

plantas existentes hacia un mayor ahorro en las mismas.<br />

• Finalmente, un nuevo apartado conteniendo los conceptos de coste, precio y beneficio<br />

obtenidos se desarrolla brevemente, para aclarar errores que normalmente se<br />

cometen en la contabilización de los costes de una instalación.<br />

La Tesis Doctoral también incluye dos partes introductorias, la primera contiene la<br />

situación en los países con escasez de agua y los métodos de desalación más comunes<br />

utilizados actualmente. La segunda parte introductoria incluye el estado actual de la<br />

teoría termoeconómica necesaria para el análisis termoeconómico de la planta.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 13


Abstract<br />

Desalination is the most important source <strong>of</strong> drinking water in arid zones, especially<br />

in the Gulf Area. Desalination consumes a lot <strong>of</strong> energy <strong>and</strong>, unfortunately, mostly<br />

from oil or natural gas. Co-generation plants providing freshwater <strong>and</strong> electricity are<br />

used in the arid areas. The combination <strong>of</strong> steam turbine plants <strong>and</strong> MSF (Multi-stage<br />

Flash) units is one <strong>of</strong> the most common schemes to meet water <strong>and</strong> energy<br />

requirements, providing almost 80% <strong>of</strong> all desalinated water in the world.<br />

A dual-purpose plant is a very complex system. Its behaviour is difficult to model,<br />

especially when all the available configurations <strong>of</strong> both sub-systems are considered.<br />

Usually plant performance is analysed separately, neglecting component interactions<br />

<strong>and</strong> possible savings from the <strong>combined</strong> systems.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> techniques are the most convenient tools to analyze these<br />

systems, because they can:<br />

• Calculate the costs <strong>of</strong> the flows <strong>and</strong> products <strong>of</strong> a plant based on physical criteria<br />

(Second Law <strong>of</strong> Thermodynamics).<br />

• Assess alternatives to save energy.<br />

• Optimize operations.<br />

• Locally optimize subsystems.<br />

• Perform energy audits <strong>and</strong> assess the fuel impact <strong>of</strong> malfunctions (operation<br />

diagnosis)<br />

This Ph. D. Thesis develops the complete thermoeconomic <strong>analysis</strong> applied in an<br />

existing steam power plant <strong>and</strong> MSF desalination unit, including cost <strong>analysis</strong>,<br />

diagnosis <strong>and</strong> local optimization <strong>of</strong> the plant. Cost <strong>analysis</strong> provides the physical<br />

costs <strong>of</strong> the main flows <strong>of</strong> the dual plant depending on operating conditions. Special<br />

emphasis was made on the interactions between the plant components <strong>of</strong> both<br />

subsystems: new concepts such as induced or intrinsic malfunction, dysfunction or<br />

the malfunction matrix were included. The results demonstrate the effect <strong>of</strong> different<br />

conditions or inefficiencies in terms <strong>of</strong> water <strong>and</strong> energy costs <strong>and</strong> additional fuel<br />

consumption during an inefficiency. Operation recommendations were also included<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


16<br />

Abstract<br />

in the <strong>analysis</strong>. Local optimization <strong>of</strong> the dual plant locates the optimum point for<br />

each operating condition <strong>and</strong> is a very powerful tool for the design <strong>analysis</strong>.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> was developed using a validated model (simulator) <strong>of</strong> the<br />

plant to determine the thermodynamic reference state at design conditions for any<br />

load point, ambient condition, operating mode etc. Plant data from a dual-plant in the<br />

Gulf were used to adapt the mathematical models. The simulator also obtained the<br />

thermodynamic state <strong>of</strong> the plant when an inefficiency is estimated in the plant<br />

diagnosis.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


CHAPTER 1<br />

Introduction<br />

Water scarcity will soon be a serious problem, especially considering the rapidly<br />

increasing world population <strong>and</strong> water consumption per capita. Fortunately, part <strong>of</strong><br />

this problem may be alleviated by desalting seawater, although this process consumes<br />

a lot <strong>of</strong> energy <strong>and</strong> may be difficult to use in non-developed countries. This Ph. D.<br />

Thesis contributes to searching for a way to reduce the energy required by desalting<br />

plants <strong>and</strong> provides tools to improve desalination technology.<br />

Several studies <strong>and</strong> international organizations focus on energy <strong>and</strong> others on water,<br />

but there seems to be a marked lack <strong>of</strong> attention on <strong>combined</strong> water <strong>and</strong> energy issues.<br />

The interaction between water production <strong>and</strong> energy is the main topic in this thesis.<br />

The main objective is to determine the validity <strong>of</strong> the thermoeconomic <strong>analysis</strong> in<br />

very complex systems like a dual-purpose power <strong>and</strong> desalination plant.<br />

This thesis considers the behavior <strong>of</strong> one <strong>of</strong> the most developed systems for providing<br />

water within the following framework:<br />

• Increasing human consumption <strong>and</strong> its consequences.<br />

• Water quality <strong>and</strong> the uses derived from its quality.<br />

• The world water crisis is mostly focused on water stressed areas. In these areas<br />

the water problem may also be solved by using desalting plants.<br />

• The interactions among the methods required to provide energy to desalt water.<br />

• The reasons for studying the steam turbine power plant + Multi-Stage Flash<br />

(MSF) desalination unit from thermodynamic point <strong>of</strong> view.<br />

• How <strong>Thermoeconomic</strong> techniques as the most useful to study complex systems.<br />

The final section <strong>of</strong> this chapter includes the structure <strong>of</strong> this Ph. D. Thesis.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


18<br />

Introduction<br />

1.1 Water requirements<br />

According to Al-Gobaisi (1999), all life depends on water <strong>and</strong> all terrestrial species,<br />

including humans, depend on fresh or non-saline water. Although the oceans<br />

represent the largest water reservoir on earth (covering three-quarters <strong>of</strong> its surface),<br />

it contains a high concentration <strong>of</strong> dissolved salts (more than the 3% <strong>of</strong> its weight).<br />

This makes it unsuitable for humans, industry <strong>and</strong> even irrigation. Less than 3% <strong>of</strong><br />

the earth's water is non-saline, <strong>and</strong> the vast majority <strong>of</strong> it is locked up in glaciers <strong>and</strong><br />

ice sheets. Water is moved around the earth in global cycles (evaporation-cloud<br />

formation-rain-percolation), but only when it is non-saline <strong>and</strong> in the liquid state, can<br />

it be used by humans. Human development <strong>and</strong> indeed civilization requires a reliable<br />

supply <strong>of</strong> even greater volumes <strong>of</strong> fresh water for drinking, cooking, washing <strong>and</strong><br />

sanitation. Furthermore, industry consumes on average 200 tons <strong>of</strong> water per ton <strong>of</strong><br />

manufactured product (Al-Gobaisi, 1997). Water also makes up more than half <strong>of</strong> the<br />

human body. An average adult drinks about 2.5 liters <strong>of</strong> water per day <strong>and</strong> needs<br />

0.75 liters a day just to stay alive. According to the World Health Organization, about<br />

150 liters <strong>of</strong> water are needed per day for a satisfactory hygienic life (Al-Gobaisi,<br />

1999). But in the South more than 1,500 million people do not have drinking water<br />

(Intermón, 1998).<br />

The imbalance between the available water resources <strong>and</strong> dem<strong>and</strong> is clear, especially<br />

in arid areas like the Arabian Gulf or Northern Africa. Human water consumption per<br />

capita in this region is very high (including domestic, agricultural <strong>and</strong> industrial uses)<br />

ranging from 300 to 1,500 liters per day. Rapidly rising incomes in some countries,<br />

with the resultant increase <strong>of</strong> living st<strong>and</strong>ards, <strong>and</strong> water losses in the network have<br />

led to even higher per capita water consumption. Intensive agriculture under arid<br />

conditions increases this dem<strong>and</strong>. The available water resources from perennial<br />

surface water, renewable ground water <strong>and</strong> reclaimed wastewater are insufficient to<br />

meet the dem<strong>and</strong>. Overexploitation <strong>of</strong> ground-water decreases ground-water levels<br />

<strong>and</strong> deteriorates water quality, including salt-water intrusion. On the basis <strong>of</strong> the past<br />

experiences in arid zones, renewable freshwater resources <strong>of</strong> 1,000 cubic meters per<br />

capita per year have been considered the limit for a chronic water scarcity that will<br />

impede development <strong>and</strong> harm human health. In terms <strong>of</strong> resources deficiency, water<br />

stress is defined as an annual renewable resource less than 1,000 cubic meters per<br />

capita per year. All the countries <strong>of</strong> the Arabian World suffer from water stress<br />

(Al-Gobaisi, 1999).<br />

1.2 Water quality <strong>and</strong> uses<br />

Water use depends on its quality. The salinity <strong>of</strong> average seawater is 34,800 ppm,<br />

although it may vary between oceans. For example, the total dissolved solids (TDS)<br />

in Arabian Gulf seawater is between 43,000 <strong>and</strong> 50,000 ppm, while the Atlantic<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


World water resources <strong>and</strong> dem<strong>and</strong><br />

Ocean has an average TDS <strong>of</strong> 36,000 ppm, <strong>and</strong> 33,600 ppm for the Pacific Ocean<br />

water (Abu Qdais, 1999).<br />

The highest limit for human consumption is 1,000 ppm (Spiegler <strong>and</strong> El-Sayed,<br />

1994), although the maximum permissible salt concentration in drinking water<br />

depends on the type <strong>of</strong> salt, the total daily water consumption <strong>and</strong> the climate (e.g., if<br />

the climate is hot <strong>and</strong> the salt is mainly sodium chloride, excess salt can even be<br />

beneficial to the human body). On average humans consume 2-8 liters per day. Thus,<br />

salt-water rejection for drinking water does not present a serious economic problem<br />

in the future, if compared with the water dem<strong>and</strong> for agricultural or industrial<br />

purposes.<br />

The purity <strong>of</strong> water for industry strongly depends on the use. Sometimes brackish<br />

water (water with less than 5,000 ppm) is enough for industrial purposes, but<br />

ultrapure water is needed for specific processes like cooling power generation plants.<br />

The amount <strong>of</strong> water for industry is several times human water consumption which<br />

is why we need more research on saving water in industrial processes <strong>and</strong> reusing<br />

waste water.<br />

Non-natural irrigation (that is, not provided by rainfall) consumes the most amount <strong>of</strong><br />

the world's water. For example, in China agriculture uses up 87% <strong>of</strong> the total water<br />

dem<strong>and</strong>. In arid areas irrigation consumes enormous amounts <strong>of</strong> water. Desalination<br />

processes are so expensive that they are not feasibly introduced to irrigate l<strong>and</strong>.<br />

However, brackish waters with a moderate salinity (about 2,000 ppm) are acceptable<br />

for some crops. The tolerance limits <strong>of</strong> each plant must be examined as a function <strong>of</strong><br />

the soil, climate, saltwater composition, irrigation method <strong>and</strong> additional treatments<br />

(fertilizers).<br />

1.3 World water resources <strong>and</strong> dem<strong>and</strong><br />

Seawater desalination is most common in the countries bordering the Persian-<br />

Arabian gulf, the north <strong>of</strong> Africa <strong>and</strong> the Canary isl<strong>and</strong>s, the Caribbean isl<strong>and</strong>s, the<br />

Pacific region (Australia, Japan, Korea <strong>and</strong> China), <strong>and</strong> the south <strong>and</strong> east <strong>of</strong> Spain, as<br />

well as various locations in the American south-west <strong>and</strong> Florida. The following is a<br />

brief explanation <strong>of</strong> water dem<strong>and</strong> <strong>and</strong> disposal in these areas in order to introduce<br />

the reader to the world’s water scarcity problem.<br />

1.3.1 Gulf Region<br />

The annual per capita annual water resources <strong>of</strong> countries in the Gulf region (United<br />

Arab Emirates, Saudi Arabia, Bahrain, Oman, Qatar <strong>and</strong> Kuwait. Iran <strong>and</strong> Iraq are<br />

excluded in the study) are very scarce. The fast growing population <strong>and</strong> increasing<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 19


20<br />

Introduction<br />

per capita water dem<strong>and</strong> (over 500 l per capita per day, Abdel-Jawad <strong>and</strong> Al-<br />

Tabtabaei, 1999) to meet the huge socio-economic developments since the 70s have<br />

recently magnified the problem. These countries are characterized by scanty rainfall<br />

<strong>and</strong> high evaporation <strong>and</strong> consumption which leads to deficits in their water budget.<br />

All these factors classify these countries as arid to semi-arid because <strong>of</strong> their limited<br />

conventional water resources <strong>and</strong> generally absent reliable surface water.<br />

Arabian Gulf seawater is quite different from other oceans:<br />

• The Arabian Gulf is roughly rectangular, surrounded by Iraq <strong>and</strong> Kuwait on the<br />

northwest, Saudi Arabia, Qatar, United Arab Emirates (UAE) <strong>and</strong> Oman on the<br />

west <strong>and</strong> south <strong>and</strong> by Iran on the east. The Gulf is approximately 100 Km long<br />

<strong>and</strong> 300 Km wide, with a surface area <strong>of</strong> 2.39×<br />

105<br />

Km2.<br />

Average water depth is<br />

35 meters, so its volume is 8.63×<br />

103<br />

Km3.<br />

Water circulates very slowly between<br />

the Arabian Gulf <strong>and</strong> the Gulf <strong>of</strong> Oman via the Strait <strong>of</strong> Hormuz: the average<br />

residence time <strong>of</strong> water is 2-5 years.<br />

• The Gulf Region has an arid sub-tropical climate with very limited annual rainfall.<br />

Water temperature varies seasonally from 18 ºC to 33 ºC. Therefore, evaporation is<br />

very high most <strong>of</strong> the year, exceeding the total river run<strong>of</strong>f by approximately a<br />

factor <strong>of</strong> 10. The effect <strong>of</strong> the river run<strong>of</strong>f, temperature <strong>and</strong> evaporation explain the<br />

gradually increasing salinity (from 36,300 to 50,000 ppm).<br />

• The Gulf ecosystem is seriously endangered <strong>and</strong> it is located in a region with<br />

political conflicts (two major wars in the last 15 years). It is also the largest oil<br />

route in the world; 20% <strong>of</strong> the total world production <strong>of</strong> oil passes through the<br />

Gulf. The serious environmental impact <strong>of</strong> large desalination units should be<br />

considered.<br />

Water stores are gradually depleting since it is extracted faster than refilled:<br />

approximately 17,000 million cubic meters are used per year <strong>and</strong> 3,000 million cubic<br />

meters recharged, <strong>and</strong> 4,000 million cubic meters are available from surface water. The<br />

total current water dem<strong>and</strong> is about 20,000 Mm3/y,<br />

with non renewable resources<br />

satisfying approx. 75% with the rest supplied by renewable conventional sources,<br />

desalination plants <strong>and</strong> recycled wastewater. Table 1.1 shows the ground water<br />

resources <strong>and</strong> the amount <strong>of</strong> renewable water resources in 1994 per year in the Gulf<br />

Countries.<br />

Table 1.1 informs that the water stress in the Gulf countries is one <strong>of</strong> the main<br />

problems that needs to be solved. Water withdrawal or water dem<strong>and</strong> is shown in<br />

table 1.2. The total dem<strong>and</strong> is divided in domestic, agricultural <strong>and</strong> industrial uses.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 1.1<br />

Country<br />

World water resources <strong>and</strong> dem<strong>and</strong><br />

Ground water disposal <strong>and</strong> renewable water resources in the Gulf Countries in 1994 (Alawadhi,<br />

1999).<br />

Population<br />

(millions)<br />

Ground water<br />

resources<br />

(Mm3/y)<br />

Conventional<br />

Renewable water resources (Mm3/y)<br />

Non conventional<br />

Desalination Wastewater<br />

Saudi Arabia 18.18 14,430 4,550 874 217<br />

UAE 2.15 1,000 490 385 110<br />

Kuwait 1.62 114 161 514 83<br />

Qatar 0.53 185 50 108 25<br />

Bahrain 0.55 190 90 75 32<br />

Oman 2.05 728 1,929 39 25<br />

Total 25.08 16,647 7,270 1,995 492<br />

TABLE 1.2<br />

Water dem<strong>and</strong> for the Gulf Countries in 1990 (ESCWA, 1994).<br />

Country<br />

Total dem<strong>and</strong><br />

(Mm3/y)<br />

Withdrawal in various sectors (Mm3/y)<br />

Domestic Agricultural Industrial<br />

Saudi Arabia 16,300 1,508 14,600 192<br />

UAE 1,490 513 950 27<br />

Kuwait 383 295 80 8<br />

Qatar 194 76 109 9<br />

Bahrain 223 86 120 17<br />

Oman 1,236 81 1,150 5<br />

Total 19,826 2,559 17,009 258<br />

Desalination is a means <strong>of</strong> augmenting fresh water resources to remove or at least<br />

reduce water stress. The number <strong>of</strong> desalination plants in the Gulf Council Countries<br />

(GCC) states increases daily. Table 1.3 summarizes the production <strong>and</strong> capacity <strong>of</strong><br />

the Middle East countries.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 21


TABLE 1.3<br />

TABLE 1.4<br />

22<br />

Introduction<br />

Total installed capacity <strong>and</strong> production in the seawater desalination plant <strong>of</strong> the Gulf Area in year<br />

1994 (Alawadi, 1999; Al-Gobaisi, 1999).<br />

Country Total capacity (m3/d)<br />

Total production (Mm3/y)<br />

Saudi Arabia 4,179,882 874.2<br />

UAE 2,066.340 385<br />

Kuwait 1,409,000 514<br />

Qatar 295,000 108<br />

Bahrain 220,571 75<br />

Oman 105,000 39<br />

Total 8,275,793 1,995<br />

Water production in Gulf countries represented the majority <strong>of</strong> the worldwide<br />

capacity. Table 1.4 shows representative values <strong>of</strong> freshwater produced in different<br />

processes. As seen in the table, large-scale Multi-stage Flash (MSF) plants installed<br />

in the Gulf produce the maximum quantity <strong>of</strong> freshwater <strong>and</strong> are the most<br />

competitive with more than 20,000 m3/d.<br />

Desalted seawater per capita per day is very<br />

high in some countries such as UAE <strong>and</strong> Qatar: 1.2 <strong>and</strong> 1.7 cubic meters per person<br />

<strong>and</strong> day.<br />

Contracted capacity <strong>of</strong> freshwater production from seawater <strong>and</strong> all waters with the existing<br />

process. The total capacity is 12.8 million cubic meters per day <strong>and</strong> 21 million cubic meters per<br />

day, respectively. Data collected in 1996 (Alawadhi, 1999).<br />

Seawater All waters<br />

World Gulf World Gulf<br />

% MSF 77.3 64.8 47.6 39.5<br />

% RO 13.3 4.7 38.6 10.9<br />

% ED — — 5.2 1.0<br />

% VC 4.6 1.5 4.3 1.0<br />

% ME 4.6 0.7 4.3 0.5<br />

Total 100 71.7 100 52.9<br />

Gulf countries actually recycle no more than 35% <strong>of</strong> their total treated wastewater,<br />

which contributes about 2.2% to the total water supply. Treated seawater is currently<br />

used mainly for l<strong>and</strong>scaping, fodder crop irrigation <strong>and</strong> some very specific industrial<br />

uses. There are a total <strong>of</strong> 105 sewage water treatment plants in the Gulf countries with<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 1.5<br />

World water resources <strong>and</strong> dem<strong>and</strong><br />

a total capacity <strong>of</strong> about 2 Mm3/d.<br />

There is no doubt that this water source is<br />

underused due to the lack <strong>of</strong> wastewater plants. More <strong>of</strong> these plants are needed make<br />

better use <strong>of</strong> this water source <strong>and</strong> minimize the serious impact on the environment as<br />

a result <strong>of</strong> its uncontrolled <strong>and</strong> unsafe disposal. Salt intrusion, ground water quality<br />

<strong>and</strong> the saline interface between sea <strong>and</strong> ground water are some <strong>of</strong> the problems that<br />

could be avoided with these plants.<br />

1.3.2 Pacific Region <strong>and</strong> India<br />

The Pacific Region is diverse in terms <strong>of</strong> desalination. Japan <strong>and</strong> Korea have<br />

developed their own desalination technology which competes on the world market.<br />

Australia <strong>and</strong> China also have their own technology <strong>and</strong> the rest <strong>of</strong> countries import<br />

plants from overseas. Here we will consider the first two categories.<br />

Table 1.5 shows the water resources in these four countries. Water resource per capita<br />

is one <strong>of</strong> the fundamental indexes <strong>of</strong> water abundance. However, they only express<br />

part <strong>of</strong> the potential availability since in some cases the transportation cost is too<br />

high. Australia, for example, has the highest water value per capita because it has a<br />

small population with rather little <strong>and</strong> irregular precipitation, <strong>and</strong> high evaporation.<br />

Japan has the most precipitation but also the largest population. In China water<br />

availability is irregular due to the climate <strong>and</strong> population distribution. Korea has the<br />

least water per capita despite <strong>of</strong> a lot <strong>of</strong> precipitation.<br />

Natural resources in the pacific region in the year 1998 (Goto et al., 1999).<br />

Country<br />

Precipitation<br />

(mm/y)<br />

Population<br />

(millions)<br />

Available water<br />

(Mm3/y)<br />

Water per capita<br />

(m3/y)<br />

Australia 465 18.1 100 5,520<br />

China 648 1,224 2,813 2,340<br />

Japan 1,714 125.6 422 3,360<br />

Korea 1,274 46.4 69.7 1,500<br />

Agricultural use occupies the largest portion in the region, whereas the consumption<br />

for living is dependent <strong>of</strong> the area (st<strong>and</strong>ard <strong>of</strong> living, life-style <strong>and</strong> climate determine<br />

the water consumption). Industrial water consumption is increased by industrial<br />

development but can be decreased by efforts such as recycling. Table 1.6 summarizes<br />

the fresh water consumption in the four countries.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 23


TABLE 1.6<br />

TABLE 1.7<br />

24<br />

Country<br />

Introduction<br />

Water use trends in the Pacific region (Goto et al., 1999).<br />

Country (year) Total (Mm3/y)<br />

% Agriculture % Living % Industry<br />

Australia (1995) 18,600 82.17 10,35 7.47<br />

China — 87 11 2<br />

Japan (1995) 90,700 58.7 17.2 14.8<br />

Korea (1996) 23,668 62.85 26.23 10.91<br />

Desalination in the Pacific region is not as important as in the Gulf region. Table 1.7<br />

explains the capacity, process, use <strong>and</strong> feed water <strong>of</strong> the desalination plants in the<br />

Pacific area.<br />

Desalination installations in the Pacific region. Data from 1998 (Goto et al., 1999).<br />

Capacity<br />

(m3/d)<br />

Australia 84,000<br />

China 182,000<br />

Japan 129,885<br />

Korea 180,000<br />

Process Use Feed water<br />

64% RO<br />

18% VC<br />

12% MSF + ME<br />

85% RO<br />

15% MSF + ME<br />

88% RO<br />

6.5% ED<br />

3.5% MSF<br />

1.8% ME<br />

> 90% RO<br />

Rest ED<br />

45% Industry<br />

33% Power gen.<br />

15% Municipal<br />

55% Industry<br />

40% Power gen.<br />

5% Living<br />

53% Industry<br />

47% Water supply systems<br />

100% Industry including<br />

power generation<br />

70% brackish<br />

18% wastewater<br />

10% seawater<br />

50% brackish<br />

20% pure water<br />

30% river, wastewater<br />

Seawater <strong>and</strong> brackish mainly<br />

Pure > brackish ><br />

wastewater > river water<br />

In conclusion, water shortage will increase with the development <strong>of</strong> industry <strong>and</strong> an<br />

improved st<strong>and</strong>ard <strong>of</strong> living in the coming century, especially in the more populated<br />

areas like China.<br />

There are more than 200,000 villages in India with inadequate drinking water, out <strong>of</strong><br />

which about 50,000 suffer from brackishness problems affecting a population <strong>of</strong><br />

about 60 million. Approximately one third <strong>of</strong> these villages are acutely affected by<br />

salinity levels above 4,000 ppm. Villages with an average population <strong>of</strong> about 500 to<br />

1,500 are mostly separated either by mountainous terrain or long stretches <strong>of</strong> barren<br />

l<strong>and</strong> <strong>and</strong> can be broadly categorized into inl<strong>and</strong> <strong>and</strong> coastal. Provision <strong>of</strong> safe<br />

drinking water to the villages inl<strong>and</strong> has been given high priority in recent years, with<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 1.8<br />

TABLE 1.9<br />

Country<br />

World water resources <strong>and</strong> dem<strong>and</strong><br />

hundreds <strong>of</strong> small Reverse Osmosis <strong>and</strong> Electrodialysis (RO/ED) plants (10-30 m3/d)<br />

installed in the affected villages. Only two Multi-Effect Distillation (MED) plants <strong>of</strong><br />

more than 10,000 m3/d<br />

were installed to supply process water in their industrial<br />

complex by seawater desalination (Prabhakar et al., 1997).<br />

1.3.3 North Africa<br />

In this region, water resources seem to be limited in time <strong>and</strong> space, unequally<br />

distributed <strong>and</strong> remote with respect to centers suffering from a continuous increase in<br />

dem<strong>and</strong>. The annual renewable water resources in this region are shown in table 1.8<br />

(Al-Gobaisi, 1997).<br />

Water disposal in the African region in 1995.<br />

Country<br />

Annual renewable water resources<br />

Total (Mm3/y)<br />

Per capita (m3/y)<br />

Algeria 14.8 528<br />

Egypt 58.1 923<br />

Libya — —<br />

Morocco 30.0 1,110<br />

Tunisia 3.9 443<br />

Water extracted from the ground is very high in some <strong>of</strong> these countries, as seen in<br />

table 1.9.<br />

Water withdrawal in North African countries. Data collected in 1990 for Algeria <strong>and</strong> Tunisia; for<br />

Egypt <strong>and</strong> Morocco data from 1992 (Al-Gobaisi, 1997).<br />

Annual withdrawal<br />

% water resources Per capita (m3/y)<br />

% Agriculture % Industrial % Living<br />

Algeria 30 180 60 15 25<br />

Egypt 97 956 85 9 6<br />

Libya — — — — —<br />

Morocco 36 427 92 3 5<br />

Tunisia 78 381 89 3 9<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 25


TABLE 1.10<br />

26<br />

Introduction<br />

In the future, desalination should be the alternative saving solution when the<br />

mobilization <strong>of</strong> non-conventional water resources is impossible or very costly<br />

(essentially in coastal zones). In this regard, five North African countries (Morocco,<br />

Algeria, Tunisia, Libya <strong>and</strong> Egypt) requested in 1989 technical assistance from the<br />

International Agency <strong>of</strong> Atomic Energy (IAAE) to study the feasibility <strong>of</strong><br />

desalination using nuclear power. The aim was to reuse the treated water in<br />

wastewater plants <strong>and</strong> provide an important resource to agriculture.<br />

There is little information about desalination plants in Northern Africa, although the<br />

water production there is almost negligible with respect to the Middle East Countries.<br />

Desalination in Egypt is the most important in the region, but the total capacity<br />

contracted is now reported to be 95,000 m3/d<br />

(Hassan <strong>and</strong> Florido, 1999). The MSF<br />

desalination technology switched to reverse osmosis for large plants over 5,000 m3/d<br />

in the last few years The proportion is 55% for the RO plants, 40% for the MSF plants<br />

<strong>and</strong> the rest in Vapor Compression (VC). Libya has two MSF plants <strong>of</strong> 24,000 <strong>and</strong><br />

10,000 m3/d<br />

(VA Tech, 1999), <strong>and</strong> in the south <strong>of</strong> Tunisia there are two brackish RO<br />

plants with a capacity <strong>of</strong> 12,000 m3/d<br />

(Cadagua, 1999). Morocco has only one RO<br />

plant with an installed capacity <strong>of</strong> more than 1,000 m3/d:<br />

the Laayoune Seawater<br />

Reverse Osmosis (SWRO) plant produces 7,000 m3/d<br />

<strong>of</strong> freshwater (NOPW, 1996).<br />

1.3.4 US experience <strong>and</strong> the Caribbean Isl<strong>and</strong>s<br />

California, Texas <strong>and</strong> Florida, the three states considered as the most arid <strong>and</strong> coastal<br />

areas <strong>of</strong> the country, will account for more than 45% percent <strong>of</strong> the nation’s total<br />

population growth between now <strong>and</strong> 2025. They are already experiencing the highest<br />

overall water deficit <strong>and</strong> droughts are also very common. As the population will<br />

continue growing in these areas, progressive approaches to meet water dem<strong>and</strong>s will<br />

be necessary (Ponce <strong>and</strong> Jankel, 1999).<br />

The total water use in the US has fallen since the 80’s since water is now used more<br />

efficiently. Table 1.10 shows the total water consumption <strong>and</strong> the use by each sector.<br />

Water use in the U.S. in 1995 (Gleick, 1998).<br />

Total use (Mm3/y)<br />

% Public % Irrigation % Thermo-industrial<br />

552,1 10.9 39.2 49.9<br />

Thermal technologies were used in the early years <strong>of</strong> desalination prior to<br />

development <strong>of</strong> RO, beginning in the 60’s with two MSF plants in Southern<br />

California <strong>and</strong> Florida. After that experience, RO technology has been successfully<br />

introduced in several plants. The use <strong>of</strong> desalination plants is steadily growing in the<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 1.11<br />

World water resources <strong>and</strong> dem<strong>and</strong><br />

US. The desalination growth rated based on increased contracted capacity was the<br />

highest in the world from 1996-1997, with about 120,000 m3/d<br />

<strong>of</strong> new freshwater.<br />

This implies a rate <strong>of</strong> growth between 10-20% per year, with a total installed capacity<br />

<strong>of</strong> more than 900,000 m3/d.<br />

(Wangnick, 1998). Much <strong>of</strong> the potable supplies utilize<br />

brackish water.<br />

Many <strong>of</strong> the isl<strong>and</strong> nations <strong>of</strong> the world are in warm sunny environments <strong>and</strong> have<br />

two significant items in common: beautiful beaches <strong>and</strong> a pernicious lack <strong>of</strong> potable<br />

water. Major economic growth is inhibited since the isl<strong>and</strong>’s population cannot<br />

enhance its agriculture <strong>and</strong> stimulate the tourist trade without a suitable <strong>and</strong><br />

consistent supply <strong>of</strong> useable water. The Caribbean sea is a good example. In Antigua,<br />

about 50% <strong>of</strong> the total drinking water requirements are supplied by a SWRO plant <strong>of</strong><br />

9,500 m3/d<br />

which substitutes an old MED plant (Barendsen <strong>and</strong> Moch, 1999). Other<br />

examples are a 10,000 m3/d<br />

SWRO plant in Nassau (Bahamas) (Andrews <strong>and</strong><br />

Shumway, 1999), a SWRO plant in Curaçao producing 9,000 m3/d<br />

<strong>and</strong> the Virgin<br />

Isl<strong>and</strong>s with 9 MED units <strong>and</strong> a <strong>combined</strong> production <strong>of</strong> 30,000 m3/d<br />

(Elovic <strong>and</strong><br />

Willocks, 1999).<br />

1.3.5 Mediterranean area <strong>and</strong> Europe<br />

Desalination in Spain started in the early 70’s in places with little water <strong>and</strong> near the<br />

coast. Here it was the only way to supplement natural water resources needed for<br />

domestic uses in highly populated isolated territories. The current <strong>and</strong> future<br />

development <strong>of</strong> the tourism industry is assured by the seawater desalination plants in<br />

those areas.<br />

The total capacity <strong>of</strong> Spanish desalination plants is now above 600,000 m3/d,<br />

<strong>and</strong> new<br />

projects for another 400,000 m3/d<br />

for urban uses are being developed <strong>and</strong> should be<br />

in operation in two years. Table 1.11 shows the seawater desalinated in Spain in 1998.<br />

Desalinated water in Spain during the year 1998 (Torres <strong>and</strong> Medina, 1999).<br />

Total (Mm3/y)<br />

% Urban & Tourism % Agriculture % Industry<br />

Seawater 95.3 94.4 5.6 —<br />

Brackish 126.57 20.4 47.6 32.0<br />

The desalination industry is located in dry Spain, that is, the southern part <strong>of</strong> the<br />

country: Balearic Isl<strong>and</strong>s, Canary Isl<strong>and</strong>s, Ceuta <strong>and</strong> the Costa del Sol. Three MSF<br />

plants were installed in Ceuta (1) <strong>and</strong> Las Palmas (2) in the 70’s, <strong>and</strong> small vapor<br />

compression units (VC) were the water supply in public delivery systems <strong>and</strong> private<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 27


TABLE 1.12<br />

28<br />

Introduction<br />

tourist resorts in the 80’s. Since then, reverse osmosis process (RO) is being used in<br />

big plants. Table 1.12 resumes the biggest desalination plants in Spain.<br />

Some <strong>of</strong> the RO desalination plants installed in Spain (Cadagua, 1999; Sánchez et al., 1997;<br />

Fayas <strong>and</strong> Novoa, 1997; Torres et al., 1999; AECYR, 1999).<br />

Plant Location Capacity (m3/d)<br />

Feed water<br />

Son Tugores Mallorca 35,000 Brackish<br />

Maspalomas Las Palmas 35,000 Brackish/Sea<br />

Marbellaa<br />

Málaga 56,000 Sea<br />

Bahía de Palma Mallorca 42,000 Sea<br />

Arrecife Lanzarote 32,500 Sea<br />

Las Palmas III Las Palmas 38,000 Sea<br />

Alicante 50,000 Sea<br />

Alicantea<br />

a. Not in operation<br />

The use <strong>of</strong> wastewater in agriculture irrigation, l<strong>and</strong>scape improvement, leisure needs<br />

<strong>and</strong> aquifer recharge is another way to supply the increasing water dem<strong>and</strong> in Spain.<br />

The Republic <strong>of</strong> Cyprus is an isl<strong>and</strong> at the eastern end <strong>of</strong> the Mediterranean Sea<br />

plagued by draught <strong>and</strong> water shortages in recent years. Seawater desalination has<br />

been the main solution. It has two little MSF plants, a MED plant <strong>and</strong> a RO plant with<br />

a capacity <strong>of</strong> 20,000 m3/d<br />

(Echaniz et al., 1997). A new RO plant with a capacity <strong>of</strong><br />

40,000 m3/d<br />

will be built by the year 2000.<br />

Desalination in the rest <strong>of</strong> Mediterranean countries is less important. There are small<br />

old MSF plants <strong>and</strong> VC units in the south <strong>of</strong> Italy to cover the local dem<strong>and</strong> (Ophir<br />

<strong>and</strong> Gendel, 1999; Italimpianti, 1999). Greece, Turkey, Jordan, Israel <strong>and</strong> Lebanon<br />

(VA Tech, 1999) also have small desalination RO plants.<br />

Germany <strong>and</strong> Austria have several desalination plants to recycle wastewater or<br />

produce pure water for industrial processes including power generation (VA Tech,<br />

1999). They do not produce drinking water.<br />

Humanity has developed non-conventional sources <strong>of</strong> potable water in order to<br />

remove or at least reduce water stress. Seawater desalination is the most important <strong>of</strong><br />

the non-conventional ways <strong>of</strong> producing water <strong>and</strong> several processes have been<br />

developed in the last few years to produce fresh water for human consumption. Yet<br />

desalinated water makes up only one part in a thous<strong>and</strong> <strong>of</strong> the fresh water used<br />

worldwide. Desalination costs several times more than conventional means <strong>and</strong> is<br />

therefore mostly used in developed countries with water scarcity (i.e. Arab countries).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 1.13<br />

Desalination <strong>and</strong> energy<br />

1.4 Desalination <strong>and</strong> energy<br />

Desalination is highly energy intensive <strong>and</strong> should not be considered in isolation<br />

from energy. The power requirements <strong>of</strong> seawater desalination plants is also<br />

increasing. There is a theoretical minimum power needed to desalt water but much<br />

more power is required in practice (El-Sayed <strong>and</strong> Silver, 1980). Unfortunately, most<br />

<strong>of</strong> the energy used is obtained from oil <strong>and</strong> natural gas. The Arab World desalinates<br />

using their large fossil fuel reserves. Consequently, the specific consumption <strong>of</strong> a<br />

desalination process must be accounted in fuel not electrical consumption as usually<br />

given when measuring plant efficiency. Table 1.13 shows the primary energy or fuel<br />

consumed in most desalination methods in the world. Note that the specific<br />

consumption has strongly decreased as desalting technology has developed.<br />

Specific consumption <strong>of</strong> desalination processes. Data obtained from several sources (Fisia-<br />

Italimpianti, 1999; I.D.E., 1999).<br />

Process MSF MED VCa<br />

Specific consumption<br />

(kJ fuel/ kg water )<br />

400-500<br />

200-300 b<br />

350-400<br />

200-250 b<br />

a. Electrical energy produced in a conventional power plant at 30% efficiency.<br />

b. Desalination process in a co-generation plant.<br />

c. Including energy recovery system in the RO process.<br />

100-200<br />

ROa<br />

70-90<br />

30-50 c<br />

As seen in the previous table, thermal distillation consumes more than other methods<br />

<strong>and</strong> more or less recovers (in the worst case) 80% <strong>of</strong> the latent heat <strong>of</strong> boiling water at<br />

atmospheric conditions (about 2,257 kJ/kg). In the previous table, specific<br />

consumption strongly depends on way the required energy is obtained (converting the<br />

primary energy from the fossil fuels into thermal or electrical energy to supply the<br />

plant). Up until recently power plant technology has developed separately from the<br />

technology used to desalt sea or brackish water. However, when the co-generation<br />

concept is applied to combine the two processes, the consumption <strong>of</strong> the desalination<br />

process can be reduced more than 50%. Including <strong>combined</strong> cycles in new MSF/<br />

MED plants considerably reduces consumption <strong>and</strong> also provides electricity in areas<br />

with energy dem<strong>and</strong>. Co-generation fuels could be substituted by biomass or refuse<br />

fuels (Tadros <strong>and</strong> Tadros, 1997). The energy-water interaction should be investigated<br />

further <strong>and</strong> improved in order to provide water to water stressed areas at minimum<br />

cost.<br />

Desalination is almost entirely powered by the combustion <strong>of</strong> fossil fuels. Their finite<br />

supply is rapidly being depleted <strong>and</strong> they also pollute the air <strong>and</strong> contribute to global<br />

climate change. Assuming that all desalinated water in the world (total installed<br />

capacity <strong>of</strong> 13 Mm3/d)<br />

is produced at an average fuel consumption <strong>of</strong> 200 kJ/kg, <strong>and</strong><br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 29


Introduction<br />

that the current annual global consumption <strong>of</strong> oil is 25 billion barrels (rising 2% per<br />

annum, Al-Gobaisi, 1997), 0.17% <strong>of</strong> world oil consumption is consumed in<br />

desalination. To underline how important energy is in desalination, if all the water<br />

consumed in the world came from desalination plants (remember that it is actually<br />

only one part in a thous<strong>and</strong>, Al-Gobaisi, 1999) the required oil would surpass the<br />

current yearly oil consumption.<br />

The development <strong>of</strong> renewable-driven desalination is still severely impeded (if not<br />

stopped) by the pressure from contemporary economic factors <strong>and</strong> political inertia. If<br />

our technology continues along the present unsustainable path, not only it is essential<br />

to have an orderly transition in the energy used for desalination (from fossil fuels to<br />

renewable resources) but the whole industry needs to gear itself towards enhanced<br />

efficiency, waste minimization <strong>and</strong> less environmental impact (Menéndez, 1997). In<br />

short, the philosophy <strong>of</strong> industrial ecology needs to be applied for desalination. The<br />

concept <strong>of</strong> industrial ecology considers an industrial system together with its<br />

surrounding systems. This systems view <strong>of</strong> industrial operations seeks to optimize the<br />

total materials cycle from raw material to manufactured material, from component to<br />

product <strong>and</strong> waste to ultimate disposal. Energy, resources <strong>and</strong> capital are the factors<br />

that have to be optimized.<br />

1.5 Why a MSF <strong>and</strong> power plant?<br />

The dem<strong>and</strong> for electricity increases every day in arid <strong>and</strong> warm areas where air<br />

conditioning is used to improve living st<strong>and</strong>ards. A dual-purpose plant is one <strong>of</strong> the<br />

best solutions to supply water <strong>and</strong> energy dem<strong>and</strong>s (although is not the most efficient<br />

method to produce fresh water, see table 1.13). As the nuclear or coal power plants<br />

are not very common in the Gulf Area, the more abundant fossil fuels like natural gas<br />

or fuel oil are consumed in new co-generation plants. Solar powered desalination is<br />

an insignificant proportion because <strong>of</strong> the costs <strong>of</strong> using renewable energy are very<br />

dependent on the scale <strong>of</strong> the infrastructure.<br />

Several power generation configurations can be coupled with a desalination unit:<br />

steam turbine plants, gas turbine plants, <strong>combined</strong> cycle power plant (gas turbine,<br />

heat recovery steam generator <strong>and</strong> steam turbine). Some desalination processes only<br />

require electrical power (not exhaust gas or steam) <strong>and</strong> co-generation is not possible.<br />

In those cases, desalination <strong>and</strong> power generation can be studied separately although<br />

the way <strong>of</strong> producing electricity is the same.<br />

This thesis aims to demonstrate the scope <strong>of</strong> <strong>Thermoeconomic</strong> Analysis when applied<br />

to a very complex system. One <strong>of</strong> the most important configurations <strong>of</strong> dual-purpose<br />

desalination plants is the multi-stage flash desalination unit (MSF) coupled with a<br />

steam turbine power plant fuelled by natural gas (fuel is also available in exceptional<br />

conditions <strong>and</strong> startups). This type <strong>of</strong> configuration is used in a plant containing the<br />

30 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Why a MSF <strong>and</strong> power plant?<br />

largest single desalination units in the world, in the United Arab Emirates (UAE).<br />

MSF units provide almost 77% <strong>of</strong> all desalinated seawater <strong>and</strong> nearly 82% <strong>of</strong> that<br />

production is from the Gulf Area (Alawadhi, 1999). MSF plants with unit capacity up<br />

to the unit studied here are likely to dominate the scene in the Gulf countries for at<br />

least another 10 years. The other predominant method <strong>of</strong> obtaining freshwater,<br />

reverse osmosis (RO), is not in favor due to the high salinity <strong>and</strong> temperatures <strong>of</strong> Gulf<br />

seawater. MSF desalination is energy intensive <strong>and</strong> inefficient especially if the steam<br />

turbine plant does not include a reheater in the boiler. It is therefore a good example<br />

to study from the thermodynamic point <strong>of</strong> view, following the Second Law<br />

perspective. The conventional energy <strong>analysis</strong> methods based on the First Law <strong>of</strong><br />

Thermodynamics are implicitly compared here.<br />

The reason for studying an MSF plant is not only its dominant position in the world<br />

desalination market. In terms <strong>of</strong> energy consumption, MSF is the worst desalination<br />

process (see table 1.13). However, from a thermodynamic point <strong>of</strong> view it <strong>of</strong>fers<br />

many more possibilities to reduce energy consumption in the process. The minimum<br />

power requirement (or thermodynamic limit) to desalt water is consumed in rejecting<br />

the difference <strong>of</strong> the equilibrium vapor pressure between saltwater <strong>and</strong> freshwater<br />

(this difference depends on the process temperature). All practical processes are<br />

non-ideal, performed by imperfect devices, <strong>and</strong> are accompanied by auxiliary nonideal<br />

processes. So, the minimum power requirement is higher for all desalination<br />

processes. In RO or VC processes, the power requirement is electrical energy<br />

produced in external power plants. Reducing the energy consumption <strong>of</strong> the process<br />

is only possible in the desalination process. But when a thermal desalination plant<br />

like a MSF unit is <strong>combined</strong> with a power plant, MSF technology can be oriented to<br />

improve the thermal efficiency <strong>of</strong> vertical tube evaporators (VTE) that allow the use<br />

<strong>of</strong> low temperature heat sources such as turbine reject steam (Sephton, 1999; Sephton<br />

<strong>and</strong> Salomon, 1997), normally rejected to the environment (through the steam cycle<br />

condenser). In the limit, the cooling tower <strong>of</strong> a conventional power plant can be<br />

substituted by a low-temperature MSF unit to highly improve the efficiency <strong>of</strong> the<br />

steam cycle. <strong>Thermoeconomic</strong> <strong>analysis</strong> connects the Second Law <strong>of</strong> Thermodynamic<br />

<strong>and</strong> Economics <strong>and</strong> is especially recommended for these two <strong>combined</strong> processes.<br />

This is the first time an in depth thermoeconomic study has been made <strong>of</strong> a<br />

desalination plant, a system combining thermal <strong>and</strong> chemical processes. Interestingly<br />

the first thermoeconomic ideas were applied to desalination processes in the sixties<br />

<strong>and</strong> early seventies (Evans, 1962; Tribus et al., 1960; Tribus <strong>and</strong> Evans, 1963;<br />

El-Sayed <strong>and</strong> Aplenc, 1970; El-Sayed <strong>and</strong> Evans, 1970), but were most developed in<br />

the eighties <strong>and</strong> nineties when <strong>Thermoeconomic</strong>s was applied in power plants.<br />

Several exergy analyses <strong>of</strong> MSF plants have already been made (Hamed et al., 1999;<br />

Darwish, Al-Najem <strong>and</strong> Al-Ahmad, 1993; Al-Sulaiman <strong>and</strong> Ismail, 1995; El-Nashar,<br />

1993), <strong>and</strong> the optimization <strong>of</strong> thermal desalting systems has also been considered<br />

(El-Sayed, 1996). In this Ph. D. Thesis, thermoeconomic techniques previously<br />

applied only to power plants were successfully used for a <strong>combined</strong> power generation<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 31


Introduction<br />

<strong>and</strong> desalination process. Chemical exergy was successfully introduced in a most<br />

complex installation, in the global exergy balance. Furthermore, no thermodynamic<br />

<strong>analysis</strong> has been done for a dual-purpose plant with two different products: water<br />

<strong>and</strong> electricity. The interactions between the two processes were analyzed in this<br />

Ph. D. Thesis. New methodologies are introduced in this complex system, allowing a<br />

better underst<strong>and</strong>ing <strong>of</strong> the real relationships between the plant equipment.<br />

1.6 <strong>Thermoeconomic</strong> <strong>analysis</strong><br />

A dual-purpose plant is a very complex system that is difficult to analyze, especially<br />

when all the available configurations <strong>of</strong> both sub-systems are considered. Usually the<br />

plants are analyzed separately, neglecting component interactions <strong>and</strong> the energy<br />

savings possible from the <strong>combined</strong> <strong>analysis</strong>. When two different products are<br />

obtained in a co-generation plant, it is very difficult to quantify the real cost <strong>of</strong> each<br />

product <strong>and</strong> redistribute the costs over the rest <strong>of</strong> upstream flows inside the dualpurpose<br />

plant by applying conventional energy <strong>analysis</strong> techniques based on the First<br />

Law <strong>of</strong> Thermodynamics.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> techniques are the most convenient tools to analyze these<br />

systems, because they can:<br />

• Calculate the costs <strong>of</strong> the flows <strong>and</strong> products <strong>of</strong> a plant based on physical criteria<br />

(Second Law <strong>of</strong> Thermodynamics).<br />

• Assess alternatives for energy savings.<br />

• Optimize operation.<br />

• Locally optimize subsystems.<br />

• Perform energy audits <strong>and</strong> assess fuel impact <strong>of</strong> malfunctions (operation<br />

diagnosis)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> uses the First <strong>and</strong> Second law <strong>of</strong> Thermodynamics in<br />

combination with economic data <strong>and</strong> introduces new concepts such as Fuel-Product,<br />

productive structure, exergy savings, cost <strong>of</strong> irreversibilities, additional fuel<br />

consumption, malfunction <strong>and</strong> others. The degradation mechanisms <strong>of</strong> the energy<br />

quality in each component require a comprehensive approach that encompasses<br />

resources, generation <strong>of</strong> products, specific unit consumption <strong>and</strong> cost, plant/system<br />

malfunction, impact on fuel consumption, etc. A better underst<strong>and</strong>ing <strong>of</strong> the actual<br />

plant performance increases the potential for improvements in operation <strong>and</strong>/or<br />

design.<br />

When applied to analyze an existing dual plant, thermoeconomic <strong>analysis</strong> requires a<br />

validated model (simulator) <strong>of</strong> the plant to determine the thermodynamic reference<br />

state at design conditions for any load point, ambient conditions, operating mode, etc.<br />

Data from a dual plant in Abu Dhabi were used to adapt the models to reproduce the<br />

32 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Ph. D. Thesis development<br />

states <strong>of</strong> the plant (therefore, the data obtained by the simulator are considered<br />

measured data). As in this case, the data acquisition, processing <strong>and</strong> storage system is<br />

not operative to be used in the thermoeconomic <strong>analysis</strong>. The simulator can obtain the<br />

thermodynamic state <strong>of</strong> the plant when an inefficiency is detected or estimated.<br />

1.7 Ph. D. Thesis development<br />

The structure <strong>of</strong> the Ph. D. Thesis is summarized as follows. First, world water<br />

resources <strong>and</strong> dem<strong>and</strong> are reviewed, especially for the Gulf area. Water quality <strong>and</strong><br />

uses are also included to inform the non-specialist readers. A brief description is then<br />

made <strong>of</strong> the most important desalination methods (Chapter 2). When the desalination<br />

unit follows a thermal principle it is usually coupled with a power generation plant.<br />

In Chapters 3 <strong>and</strong> 4 the mathematical models applied to the power <strong>and</strong> desalination<br />

plant are developed. The results are compared <strong>and</strong> readapted with operational data<br />

from the data acquisition system <strong>of</strong> the plant. The mathematical model was validated<br />

as a tool that widely reproduces the real state <strong>of</strong> the plant under different operating<br />

conditions, as if the results were real plant data. An interactive steady-state simulator<br />

was made that can be used on a personal computer to help obtain output data.<br />

The simulator (Chapter 5) supplied the main part <strong>of</strong> this Ph. D. Thesis: the complete<br />

thermoeconomic <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant (Chapter 7).<br />

After explaining the fundamental concepts <strong>of</strong> <strong>Thermoeconomic</strong>s (Chapter 6), the first<br />

step was to build the thermoeconomic model. The most convenient productive<br />

structure was chosen for the power <strong>and</strong> desalination plant. The thermodynamic<br />

operation <strong>and</strong> economic costs <strong>of</strong> every flowstream <strong>of</strong> the plant were calculated <strong>and</strong><br />

analyzed. Those costs allow cost assessment <strong>of</strong> the plant products based on physical<br />

criteria. Then, the thermoeconomic diagnosis was applied. The steady-state diagnosis<br />

<strong>of</strong> the dual-purpose plant helped us obtain a more cost-effective operation <strong>and</strong> a better<br />

underst<strong>and</strong>ing <strong>of</strong> plant performance. The mathematical model was applied for a given<br />

operating condition characterized by operational data (previously validated <strong>and</strong><br />

processed) to quantitatively analyze the following steps:<br />

• Comparison with a reference case (target) with the same operating conditions.<br />

• Identification <strong>of</strong> inefficiencies, <strong>and</strong> the performance degradation <strong>of</strong> sub-systems<br />

or components. These inefficiencies were simulated.<br />

• Evaluation <strong>of</strong> the causes <strong>of</strong> cost generation <strong>and</strong> component inefficiencies.<br />

• Assessment <strong>of</strong> the extra-operating cost due to malfunctions with respect to the<br />

most feasible operation <strong>and</strong> the cost impact <strong>of</strong> appropriate maintenance actions.<br />

The previous cost <strong>analysis</strong> is therefore essential to perform the diagnosis <strong>of</strong> the<br />

system.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 33


Introduction<br />

• Operation recommendations for the plant managers, taking into account the<br />

experience from the <strong>analysis</strong> (assessment <strong>of</strong> alternatives).<br />

A new method is introduced to develop the thermoeconomic diagnosis, including the<br />

matrix formulation <strong>and</strong> some new concepts like induced <strong>and</strong> intrinsic malfunction,<br />

<strong>and</strong> dysfunction.<br />

Once the diagnosis was completed, a global optimization <strong>of</strong> the plant was performed<br />

from locally optimizing the system units. The local optimization <strong>of</strong> a unit consists in<br />

finding the minimum cost <strong>of</strong> the product <strong>of</strong> each component. The thermoeconomic<br />

model was also used in this process.<br />

Finally, the idea <strong>of</strong> maximum benefit in water <strong>and</strong> electricity production was<br />

analyzed using practical examples. The contribution <strong>of</strong> the price policy applied in the<br />

final benefit is considered by separating the methods <strong>of</strong> assessing product price <strong>and</strong><br />

cost.<br />

The last chapter (Chapter 8) contains the conclusions <strong>of</strong> the Ph. D. Thesis <strong>and</strong> future<br />

lines <strong>of</strong> research.<br />

34 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


CHAPTER 2<br />

Desalination processes<br />

In chapter 1, the great problem <strong>of</strong> water scarcity <strong>and</strong> desalination as the way to solve<br />

it is remarked. Desalination is the process that convert brackish or seawater in water<br />

for human consumption, there are several processes technologically developed<br />

providing water in arid areas.<br />

This chapter includes a general review <strong>of</strong> desalination methods, in order to have an<br />

overall perspective <strong>of</strong> the state <strong>of</strong> the art in desalination technology. The importance<br />

<strong>of</strong> the MSF with respect to the other methods is also argued in this chapter.<br />

The most reliable techniques <strong>of</strong> seawater desalination are rated into three categories<br />

depending on the principle applied:<br />

• Processes involving a change <strong>of</strong> phase: Freezing or distillation.<br />

• Processes using membranes: Reverse osmosis or electrodialysis.<br />

• Processes acting on chemical bonds: Ion exchange.<br />

Among the processes above, distillation <strong>and</strong> reverse osmosis processes show high<br />

performances in seawater desalination; thus they are the most marketable in the<br />

world. Next, we develop the following processes in detail:<br />

• Multi-Stage Flash (MSF).<br />

• Multi-Effect Distillation (MED).<br />

• Reverse Osmosis (RO).<br />

• Vapor Compression (VC).<br />

We also mentioned the other techniques, which have not been developed in the field<br />

<strong>of</strong> desalination due to problems generally, related to energy consumption <strong>and</strong>/or to the<br />

high investments required. These techniques are:<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Desalination processes<br />

• Solar Distillation.<br />

• Freezing.<br />

• Electrodialysis.<br />

• Ion Exchange.<br />

2.1 Phase change processes: distillation <strong>and</strong> freezing<br />

More than 85% <strong>of</strong> the world’s desalted water is obtained by distillation. Desalination<br />

by distillation involves boiling water seawater to release water vapor <strong>and</strong> dissolved<br />

gasses, leaving behind the salts (which are only volatile above 300 ºC). Pure water is<br />

collected by condensing the vapor inside or on the outside <strong>of</strong> tubes which may be<br />

arranged horizontally or vertically depending on the installation. Every distillation<br />

system must also be ventilated to extract air <strong>and</strong> non-condensable gases in the<br />

seawater, <strong>and</strong> a vacuum pump or steam ejector is required when the evaporatorcondenser<br />

system is at lower than atmospheric pressure.<br />

2.1.1 Multi-stage flash process (MSF)<br />

Multi-Stage Flash is the most widely used evaporation process (Wangnick, 1998).<br />

It is especially common wherever the temperature, salt content, biological activity or<br />

pollution level <strong>of</strong> raw water is high, as in the Middle East. MSF also be used if the<br />

desalination plant is coupled to a power station or if waste heat is present (e.g. from<br />

gas turbine effluents). In general, MSF plants are more common because they are<br />

simple <strong>and</strong> robust, although their specific consumption may be higher than other<br />

3<br />

methods (12-24 kWh/m ).<br />

Flash evaporation takes place when a fluid is heated to a certain temperature <strong>and</strong><br />

evaporates both above <strong>and</strong> below the atmospheric pressure: under gradual decreasing<br />

pressure, flashing by pressure reduction is called flash evaporation. In multi-stage<br />

flash plants seawater (pumped through heat exchanger tubes installed in the various<br />

evaporator stages) is heated to a certain temperature. Final heating is performed by<br />

steam in a final heater. The hot seawater then goes into flash chambers where the<br />

pressure is maintained below the equilibrium pressure corresponding to the<br />

temperature at which the brine enters. Part <strong>of</strong> the brine flashes into vapor <strong>and</strong> after<br />

passing a demister, it condenses outside the tubes while heating the seawater flowing<br />

through the tubes. The multi-flash distillation unit contains cells assembled in series,<br />

at a different pressure. The water produced in each stage is collected in a trough<br />

mounted below the tube bundle which collects the fresh water end product. These<br />

widely used units perform recycle brine (50% to 70% <strong>of</strong> the brine quantity within the<br />

last stage is collected <strong>and</strong> discharged through the seawater feeding pipe <strong>of</strong> the unit) in<br />

order to reduce the quantity <strong>of</strong> the make-up seawater needed to produce fresh water.<br />

The concentrated seawater is also removed from the last stage by a pump or by<br />

gravity. Figure 2.1 shows a general scheme <strong>of</strong> a conventional MSF unit.<br />

36 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 2.1<br />

Phase change processes: distillation <strong>and</strong> freezing<br />

General outlay <strong>of</strong> MSF distillation with brine recycling.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

37


Desalination processes<br />

Seawater with 40,000 to 50,000 ppm dissolved solids is converted into distillate <strong>and</strong><br />

fresh water with a few ppm <strong>of</strong> solids. An MSF type plant operates between two<br />

temperatures: the top brine temperature (brine heater outlet temperature, or TBT) <strong>and</strong><br />

the last stage temperature. The top brine temperature depends on:<br />

a) Available steam quality.<br />

b) Scale prevention technique.<br />

c) Brine concentration <strong>and</strong> nature <strong>of</strong> dissolved salts<br />

The last stage temperature depends on:<br />

a) Cooling water inlet temperature.<br />

b) Absolute pressure maintained in the last stage by the ejector system.<br />

In practice, MSF plants are designed for various gain outputs ratios (GOR, tons <strong>of</strong><br />

fresh water produced per tons <strong>of</strong> steam supplied to the brine heater). In practice, a<br />

G.O.R <strong>of</strong> 12:1 being the upper limit. Obviously, the production rate is a direct<br />

function <strong>of</strong> the flashing brine flow <strong>and</strong> the flash range (brine top temperature-last<br />

stage temperature). Also, in theory, the actual number <strong>of</strong> stages is not important for a<br />

given ratio.<br />

However, the number <strong>of</strong> stages determines the total exchange area required for heat<br />

recuperation. More stages will decrease the total exchange area required thereby<br />

limiting the maximum number <strong>of</strong> stages per plant. In practice, however, stage number<br />

increases at increasing gain ratios but also depends on the plant’s capacity. The<br />

number <strong>of</strong> stages is generally about 20 <strong>and</strong> sized to keep the temperature difference<br />

constant between stages (the temperature difference is estimated to be about 3 ºC).<br />

2.1.2 Multi-effect distillation (MED)<br />

Contrary to MSF, in Multi-Effect Distillation (MED) evaporation takes place on<br />

surfaces, by exchanging the latent heat through the heat transfer surface between<br />

condensing vapor on one side <strong>and</strong> evaporating brine on the other. The MED plant<br />

also has several stages, each with a heat exchanger tube bundle (see fig. 2.2).<br />

Seawater is sprayed onto the tubes <strong>and</strong> the condensing heating steam inside the<br />

tubes evaporates part <strong>of</strong> the seawater on the outside. The steam produced is used as<br />

heating steam in the next stage, where it condenses inside the tubes. The condensate<br />

is the water product. Obviously the boiling temperatures (<strong>and</strong> pressures) in the<br />

different evaporators cannot be the same. The specific consumption depends on the<br />

steam conditions supplied to the first stage, but is usually lower than in MSF<br />

3<br />

(10-15 kWh/m ).<br />

38 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 2.2<br />

Phase change processes: distillation <strong>and</strong> freezing<br />

Flow diagram <strong>of</strong> Multi-Effect Distillation (MED) with thermal vapor compression (TVC).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

39


FIGURE 2.3<br />

Desalination processes<br />

MED process with vertical tube evaporators (VTE).<br />

40 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Phase change processes: distillation <strong>and</strong> freezing<br />

The first stage is heated by external steam from a heat recovery system or a<br />

back-pressure steam turbine. But in most cases, MED plants are equipped with<br />

thermal vapor compressors for better efficiency. A steam ejector driven by mediumpressure<br />

steam removes a part <strong>of</strong> the steam produced in the last stage <strong>and</strong> compresses<br />

it to use as the heating steam. The steam produced in the last stage is condensed on<br />

the outside <strong>of</strong> exchanger tubes in a separate condenser, which is cooled by incoming<br />

seawater. Part <strong>of</strong> the heated seawater is then used as feedwater. Product water <strong>and</strong><br />

concentrated seawater are then pumped out from the last stage <strong>of</strong> the evaporator.<br />

Most MED plants have horizontal evaporators. Vertical tube evaporators (VTE) are<br />

also available: In vertical tube evaporation, salt water falls in a thin film through<br />

vertical tubes in a large chamber (figure 2.3). As it falls, it is heated by steam that<br />

condenses on the outer surface <strong>of</strong> the tubes. This heat exchange converts some <strong>of</strong> the<br />

salt water in the tubes into steam <strong>and</strong> some <strong>of</strong> the steam around the tubes into fresh<br />

water (condensate).<br />

Steam generated inside the tubes in the first chamber flows to the second chamber,<br />

<strong>and</strong> condenses on the tubes there. The process is repeated in several chambers <strong>and</strong> is<br />

sometimes called “multiple-effect falling-film” distillation, because each bundle <strong>of</strong><br />

tubes is an “effect”, <strong>and</strong> because a thin film <strong>of</strong> water falls down the inside surface <strong>of</strong><br />

the tubes. Vertical tube evaporators are most cost-effective in large plants requiring<br />

high efficiency. They have an improvement over older systems since less heat transfer<br />

surface is required <strong>and</strong> the water need only be circulated once.<br />

2.1.3 Vapor compression (VC)<br />

Thermocompression (TVC) or vapor compression distillation (VC) involves boiling a<br />

liquid (seawater in this case) on one side <strong>of</strong> the heat transfer surface, <strong>and</strong> directing the<br />

compressed vapor to the other side <strong>of</strong> the heat transfer surface to be condensed (see<br />

flow diagram, figure 2.4).<br />

In the specific design described here as an example, a single-stage VTE type seawater<br />

is boiled inside a bank <strong>of</strong> enhanced surface tubes. The generated vapor then passes<br />

through a mist separator to remove any entrained salt-water droplets. In a vertical<br />

tube evaporator, the pure vapor enters the compressor at 101,5 ºC <strong>and</strong> 1 psig for a<br />

compressed steam temperature <strong>of</strong> 106 ºC <strong>and</strong> 3.6 psig (the pressure is therefore<br />

increased 0.18 bar). The compressor is a centrifugal, single-stage type designed for<br />

high-volumetric flows. This higher-energy compressed steam is discharged into the<br />

evaporator onto the outside <strong>of</strong> the enhanced surface tubes, where it condenses <strong>and</strong><br />

provide its latent heat energy to the boiling seawater inside the tubes.<br />

Note that the process is very efficient thermodynamically, because most <strong>of</strong> the shaft<br />

work required by the compressor is used to avoid the boiling point elevation <strong>of</strong><br />

seaweater (BPE). Additional vapor is generated <strong>and</strong> the process continues. The<br />

vapor, which condenses on the outside <strong>of</strong> the tubes, is collected, <strong>and</strong> drawn <strong>of</strong>f by<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

41


FIGURE 2.4<br />

Desalination processes<br />

the distillate pump <strong>and</strong> pumped through a three-stream heat exchanger. The excess<br />

feed water, called blowdown, which is concentrated, is also pumped through the<br />

same heat exchanger. The distillate <strong>and</strong> blowdown are cooled therein while<br />

preheating the incoming feedwater. This heat exchanger helps to minimize energy<br />

consumption <strong>of</strong> the system, in a VC system the specific electric consumption is<br />

3<br />

lower than 10 kWh/m .<br />

Flow diagram <strong>of</strong> a vapor compression system with vertical tube evaporators (VTE).<br />

42 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Phase change processes: distillation <strong>and</strong> freezing<br />

Distilled water is made by condensing above atmospheric pressure at 106 ºC. A small<br />

amount <strong>of</strong> make-up heat is required for continuous operation to replace the heat lost<br />

to radiation <strong>and</strong> venting <strong>and</strong> the portion not reclaimed in the three-stream heat<br />

exchanger. Electric immersion heaters, a steam coil, or heat recovery exchangers to<br />

recover waste heat from engine jacket water or exhaust gas when available can<br />

provide this make-up heat. The distillate must be sterilized to meet Health Service<br />

requirements <strong>and</strong> may also be chlorinated for storage purposes.<br />

2.1.4 Solar distillation<br />

Solar stills use can be an ideal source <strong>of</strong> fresh water for drinking <strong>and</strong> agriculture in<br />

arid, isolated zones. Solar energy has a definite advantage over fossil energy, for<br />

small st<strong>and</strong>-alone units in rural <strong>and</strong> isolated areas (India). However, solar distillation<br />

is not widely used since installation costs are high <strong>and</strong> only a few liters can be<br />

produced per day, per square meter <strong>of</strong> pan area in the stills. Of course any economic<br />

or energetic comparison should not be considered.<br />

Several different configurations can be used to recycle the recuperated heat from the<br />

vapor condensation in solar stills. But we will only consider the conventional solar<br />

still (figure 2.5). The sun heats salt water in a black pan covered with a sloping glass<br />

ro<strong>of</strong>. Water vapor rises to the glass where it condenses, forming a film which runs <strong>of</strong>f<br />

into a collecting trough <strong>and</strong> is stored. The water does not boil but vaporizes slowly<br />

through a layer <strong>of</strong> water-saturated air <strong>and</strong> reaches the cooler glass by convection. The<br />

rate <strong>of</strong> evaporation is primarily controlled by the intensity <strong>of</strong> the incoming solar<br />

radiation which creates both temperature <strong>and</strong> water vapor concentration differences<br />

between the water <strong>and</strong> glass surface. Additional solar radiation can be obtained using<br />

lenses, mirrors <strong>and</strong> other focusing devices, but also heat losses increase when the<br />

temperature inside the solar still change. Finally, wind velocity has a negative effect<br />

on the cooling <strong>of</strong> the heating surface.<br />

The principle <strong>of</strong> the thermal energy extraction from a solar pond or other methods<br />

could be used as the energy source for seawater desalination processes. For example,<br />

the use <strong>of</strong> parabolic trough collectors (PTC) could make competitive the use <strong>of</strong> solar<br />

energy for desalination processes (MSF, García <strong>and</strong> Gómez, 1999; MED, García,<br />

Palmero <strong>and</strong> Gómez, 1999), depending on conventional energy costs, the solar<br />

collectors cost <strong>and</strong> the climatic conditions that determine the attainable fresh water<br />

2<br />

3<br />

production per m <strong>of</strong> solar collector (the PTC collectors provide on average 10 m <strong>of</strong><br />

2<br />

fresh water per m <strong>of</strong> solar collector), <strong>and</strong> the solar fraction SF that determines the<br />

percentage <strong>of</strong> the day in which the desalination plant consumes solar energy.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

43


FIGURE 2.5<br />

Desalination processes<br />

Diagram model <strong>of</strong> a solar still.<br />

Solar energy<br />

2.1.5 Freezing process<br />

Glass<br />

Condensed vapor<br />

Vapor<br />

Salt water<br />

Insulation<br />

Distilled water Distilled water<br />

This process, also based on phase change, is independent <strong>of</strong> the water salt content.<br />

Seawater is cooled <strong>and</strong> the ice is collected (ice crystals are essentially salt free). Ice<br />

formation is analogous to distillation in this respect since salt-free vapor is produced<br />

while the liquid may have a high salt concentration. The ice is melted to obtain fresh<br />

water (the fusion temperature is less than that <strong>of</strong> salts contained in the ice).<br />

The freezing process is different from distillation since the latter is carried out well<br />

above ambient temperature <strong>and</strong> the equipment is designed for minimal heat losses.<br />

In freezing methods, the system must be protected against heat gains or cold losses,<br />

<strong>and</strong> ice needs to be transported <strong>and</strong> purified, which is somewhat more complex than<br />

h<strong>and</strong>ling fluids alone. Although the low operating temperature <strong>of</strong> freezing processes<br />

greatly reduces scale <strong>and</strong> corrosion problems, refrigeration technology may be<br />

adapted. So that water is the first or secondary refrigerant. This secondary refrigerant<br />

system could be mixed or separated from water by a heat transfer surface.<br />

Freezing methods are not widely used in the desalination industry, <strong>and</strong> to calculate<br />

their power consumption, we still have to rely on experiments in relatively small <strong>and</strong><br />

medium-sized plants <strong>and</strong> extrapolation to larger plants.<br />

44 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Processes using membranes<br />

2.2 Processes using membranes<br />

2.2.1 Reverse osmosis<br />

Osmosis is a physical process which occurs naturally in animals <strong>and</strong> plants<br />

(figure 2.6). Osmotic pressure is measured using a recipient divided into 2<br />

compartments by a semi-impermeable membrane. Saline solution is poured into one<br />

half <strong>and</strong> freshwater into the other. Part <strong>of</strong> the fresh water will flow through the<br />

membrane into the saline solution. The excess height at the saline solution over the<br />

fresh water is a measure <strong>of</strong> the osmotic pressure <strong>of</strong> the solution.<br />

If external pressure greater than osmotic pressure is applied to the saline solution, the<br />

water will flow through the membrane in the other direction, leaving behind a more<br />

highly concentrated salt solution. This process is called reverse osmosis (RO). The<br />

osmotic pressure <strong>of</strong> a solution is directly proportional to the solute concentration, <strong>and</strong><br />

the permeated water flow is proportional to the difference between the applied<br />

pressure <strong>and</strong> the osmotic pressure <strong>of</strong> the concentrated solution.<br />

RO can be used to demineralize brackish water with 1-10 gr/l salinity. It is also used<br />

for seawater desalination <strong>and</strong> has lower energy consumption, investment cost, space<br />

requirements <strong>and</strong> maintenance than other processes. However, RO seawater plants in<br />

the Gulf Area need an intensive water pre-treatment process with a lower product<br />

quality, <strong>and</strong> are not <strong>of</strong>ten used.<br />

In RO desalination (figure 2.7) seawater is pretreated to avoid membrane fouling.<br />

It then passes through filter cartridges (a safety device) <strong>and</strong> is sent by a high-pressure<br />

pump through the membrane modules (permeators). Because <strong>of</strong> the high pressure,<br />

pure water permeates through the membranes <strong>and</strong> the seawater is concentrated. The<br />

water product flows directly from the permeators into a storage tank, <strong>and</strong> the<br />

concentrated seawater (at high pressure) is sent via an energy recovery system back<br />

into the sea. The four main parts <strong>of</strong> the RO installation are:<br />

Preliminary treatment unit<br />

The treatment has the following steps:<br />

•<br />

•<br />

•<br />

•<br />

•<br />

Chlorination:<br />

To reduce bacteriological <strong>and</strong> organic loads found in raw water.<br />

Filtration on a s<strong>and</strong> bed:<br />

To reduce raw water turbidity.<br />

Acidification:<br />

Acid is added to clarified raw water to lower its pH <strong>and</strong> limit the<br />

formation <strong>of</strong> calcareous deposits.<br />

Inhibition by polyphosphates:<br />

Polyphosphates delay the formation <strong>of</strong> precipitates<br />

such as calcium <strong>and</strong> barium sulfate.<br />

Dechlorination:<br />

To remove the residual chlorine from the pre-treatment.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

45


FIGURE 2.6<br />

Desalination processes<br />

Reverse osmosis process.<br />

46 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 2.7<br />

Processes using membranes<br />

Reverse osmosis (RO) desalination with Pelton turbine.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

47


Desalination processes<br />

•<br />

Cartridge filtering:<br />

To catch the particles obtained by oxidation <strong>of</strong> dissolved ions<br />

(Fe++) in raw water.<br />

Note that distillation methods only need a light chlorination process <strong>and</strong> some scale<br />

inhibitors (addition <strong>of</strong> polyphosphates). Sometimes acid is added to prevent the<br />

scaling problem.<br />

High-pressure pumping system<br />

This stage is the least problematic <strong>and</strong> normally involves centrifugal pumps.<br />

RO modules<br />

The main modules used for RO seawater desalination are made out <strong>of</strong> hollow fibers<br />

<strong>and</strong> spiral fibers provided by several manufacturers. The spiral-wound <strong>and</strong> hollow<br />

fiber designs were developed to contain the high-pressure fluid in the lowest possible<br />

volume for a given membrane surface.<br />

In spiral-wound elements membranes <strong>and</strong> backing are wound similar to a jelly roll<br />

around a central perforated tube which collects the product. Saline water flows<br />

through separate channels in one direction; the membrane elements are typically<br />

30-120 cm long <strong>and</strong> 10-30 cm in diameter. They can be mounted in series with antitelescoping<br />

devices between adjacent elements to form modules. Separate modules<br />

can readily be connected in series or in parallel.<br />

The hollow fiber units have a very large number <strong>of</strong> hollow fibers, thinner than human<br />

air, with their ends potted in epoxy resin, are held in a pressure vessel. Pressurized<br />

saline water circulates on the outside <strong>of</strong> the fibers while the hyperfiltrate flows within<br />

the fibers toward the open ends <strong>of</strong> the fibers held in position by the epoxy resin.<br />

Desalted water emerging from millions <strong>of</strong> open fiber ends is collected there. The<br />

hollow fibers are made by methods similar to those developed in the textile fiber<br />

industry. These units pack more membrane surface per unit volume than spiralwound<br />

unit <strong>and</strong> are extensively used for seawater RO.<br />

The brine energy recovery system<br />

In the last years, investigators have tried to reduce the energy requirements<br />

3<br />

(6-8 kWh/m ) <strong>of</strong> RO seawater desalination using two main devices:<br />

•<br />

•<br />

Pelton turbines:<br />

The high-pressure concentrate from membranes pushes on the<br />

Pelton blades to provoke a pair in a common shaft. Energy recovery for RO plants<br />

results in energy savings <strong>of</strong> 40% (Calder, 1999).<br />

Pressure exchangers (PE)<br />

: The PE unit uses the principle <strong>of</strong> positive<br />

displacement to pressurize low-pressure raw seawater by direct contact with the<br />

concentrate stream from a seawater membrane system. A cylindrical rotor with<br />

longitudinal ducts parallel to its rotational axis is used to transfer the pressure<br />

48 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Processes acting on chemical bounds<br />

energy from the concentrate stream to the feed stream. The energy recovery with<br />

PE is in the range <strong>of</strong> 50-65% (Hauge <strong>and</strong> Ludvigsen, 1999).<br />

2.2.2 Electrodialysis (ED)<br />

This process is used to demineralize brackish water by making different ions<br />

migrate through selective membranes in electric field made by the dirct difference <strong>of</strong><br />

voltage potential between two electrodes connected at the boundaries <strong>of</strong> the<br />

membranes.<br />

Whenever salt water is flowing in a cell, the cations are attracted by the anode <strong>and</strong> the<br />

anions by the cathode. If not constrained, these ions discharge on the electrodes <strong>of</strong><br />

opposite sign. In return, if a set <strong>of</strong> selective <strong>and</strong> permeable membranes is placed<br />

between the electrodes, salt concentration decreases in some compartments <strong>of</strong> the cell<br />

where water is desalinated, while this concentration increases in the other<br />

compartments where salt water becomes even more concentrated. This process<br />

(shown in fig. 2.8) is suitable for desalinating brackish waters with an average salt<br />

3<br />

content between 1 to 3 g/l with a very low power consumption (about 1 kWh/m ) <strong>and</strong><br />

a salt rejection <strong>of</strong> 75% (data obtained from De Armas, Torrent <strong>and</strong> Von Gottberg,<br />

1999). Above this it becomes more costly than competitive processes (its energy<br />

consumption for seawater desalination is much higher).<br />

2.3 Processes acting on chemical bounds<br />

2.3.1 Ion exchange<br />

Ion-exchanging resins are insoluble substances. In contact with a solution, they<br />

exchange some ions with the dissolved salt.<br />

Two types <strong>of</strong> resins can be used: anionic resins that substitute water anions by<br />

OH-- ions (hydroxil permutation); <strong>and</strong> cationic resins substitute cations by H+ ions<br />

(acidic permutation).<br />

Ion exchange demineralization provides high purity water if the salt concentration<br />

does not exceed 1 g/l. It is <strong>of</strong>ten used for water preparation <strong>of</strong> boilers from water <strong>of</strong><br />

streams or aquifers, characterized by their low salt content, <strong>and</strong> for s<strong>of</strong>tening water<br />

with excessive calcium <strong>and</strong> magnesium. Resins must be regenerated regularly with<br />

chemical reagents to substitute its original ions <strong>and</strong> those fixed by the resin.<br />

The resins <strong>and</strong> chemicals must be substituted regularly, which raises the cost <strong>and</strong><br />

makes it unpractical for seawater desalination.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

49


FIGURE 2.8<br />

Desalination processes<br />

Electrodialysis process.<br />

50 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Summary<br />

2.4 Summary<br />

A general review <strong>of</strong> desalination technology has been written in this chapter. The<br />

review includes the principle <strong>of</strong> operation, description <strong>of</strong> the necessary installation,<br />

advantages/disadvantages, characteristic parameters (including specific consumption)<br />

<strong>and</strong> application range <strong>of</strong> each desalination method that is now available in<br />

desalination market.<br />

MSF is not only the most dominant process in desalination. It <strong>of</strong>fers the possibility to<br />

be connected to several heat sources: steam turbines, gas turbines, solar storage,<br />

<strong>combined</strong> cycles. So, it allows applying techniques oriented to produce the MSF<br />

product with the lowest cost. This Ph. D. Thesis develops one <strong>of</strong> those techniques,<br />

nd based on 2 Law <strong>of</strong> Thermodynamics.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

51


CHAPTER 3<br />

MSF desalination<br />

steady-state model<br />

The daily world production <strong>of</strong> drinkable water from Multi-Stage Flash plants (MSF)<br />

far exceeds that <strong>of</strong> other desalination methods. This is particularly the case where<br />

power generation is linked to water production to use the process steam.<br />

In this chapter I will describe a mathematical model used in the SIMTAW program, to<br />

simulate a MSF plant under different operating conditions.<br />

MSF plant design data were included in the mathematical model, which is not<br />

oriented for design <strong>analysis</strong>. Several operating variables can be modified by the user<br />

to observe changes in plant behavior, such as consumed steam, inlet water<br />

temperature, water mass flow rates, TBT value, fouling factors <strong>and</strong> more variables<br />

explained below. The inverse calculation procedure option can evaluate the fouling<br />

factor <strong>of</strong> the stages instead <strong>of</strong> the distillate temperature pr<strong>of</strong>ile.<br />

This model provides information to perform the exergy <strong>and</strong> thermoeconomic <strong>analysis</strong><br />

<strong>of</strong> the whole dual-purpose plant, i.e. power generation plant <strong>and</strong> MSF plant, in order<br />

to analyze plant efficiency <strong>and</strong> cost savings.<br />

The structure <strong>of</strong> this section is as follows:<br />

• First, brief descriptions <strong>of</strong> the physical processes in a MSF plant.<br />

• Second, an explanation <strong>of</strong> the mathematical model, including the equations used<br />

to solve the model.<br />

• Third, a description <strong>of</strong> the solution algorithm <strong>of</strong> the system <strong>of</strong> model equations.<br />

• Finally an explanation <strong>of</strong> the <strong>simulation</strong> options <strong>and</strong> the design data.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 3.1<br />

54<br />

MSF desalination steady-state model<br />

3.1 Process description<br />

Many multi-stage flash plant arrangements <strong>and</strong> operational techniques are available.<br />

Each evaporator is usually described by defining the three main plant characteristics:<br />

the flashing flow system, the chemical treatment <strong>and</strong> the tube configuration. The MSF<br />

Plant studied here is a brine recirculation flow, high-temperature (HT) antiscale<br />

treatment, <strong>and</strong> cross tube configuration, the most typical <strong>of</strong> the MSF plant types. It<br />

3<br />

has six 20-stage condensing lines which deliver up to 14,400 m /h <strong>of</strong> water with a<br />

steam turbine cycle to provide electrical power.<br />

The plant has a single effect MSF evaporator with recycled brine (see figure 3.1).<br />

Recycled brine plants contain three main sections from left to right: the ‘heat input<br />

section’ (or brine heater), the ‘heat recovery section’ , <strong>and</strong> the ‘heat rejection section’ .<br />

The recovery <strong>and</strong> rejection sections both have a series <strong>of</strong> stages. Each stage has a<br />

flash chamber <strong>and</strong> a heat exchanger/condenser, where vapor (flashed <strong>of</strong>f in the flash<br />

chamber) is condensed. The flash chamber is separated from the condenser by a<br />

demister, where entrained brine droplets are removed from the flashing vapor, <strong>and</strong> a<br />

distillate trough catches the condensate from the condenser above.<br />

Schematic diagram <strong>of</strong> a single effect MSF evaporator with recycled brine.<br />

A brief description <strong>of</strong> the MSF desalination flow process follows (see figure 3.1). The<br />

plant feed, SR, is allowed to pass through the heat rejection section, which rejects the<br />

excess thermal energy from the plant <strong>and</strong> cools the product <strong>and</strong> brine to the lowest<br />

possible temperature when it comes from the last recovery section stage.<br />

At the output <strong>of</strong> the first (warmest) rejection stage the feed stream splits into two parts,<br />

reject seawater CW (which is returned back to the sea) <strong>and</strong> a make up stream F (which<br />

is then <strong>combined</strong> with the recycle stream). The <strong>combined</strong> stream R passes through the<br />

heat exchangers <strong>of</strong> the recovery section, where its temperature increases as it proceeds<br />

towards the heat input section <strong>of</strong> the plant. In the brine heater, the brine temperature is<br />

raised from TF,1<br />

to a maximum value TB,o<br />

( = Top Brine Temperature TBT)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 3.2<br />

Process description<br />

approximately equal to the saturation temperature at the system pressure. If the<br />

seawater temperature is lower than 25 ºC, the temper system takes part <strong>of</strong> the cooling<br />

reject seawater, so that the distiller feed temperature is at least the above mentioned<br />

temperature.<br />

The brine then enters the first heat recovery stage where it is flashed by reducing the<br />

pressure in a throttling valve. As the brine was already at its saturation temperature<br />

for a higher pressure, superheated water vapor is generated in the throttling process.<br />

This vapor passes through a wire mesh (demister), to remove any entrained brine<br />

droplets before condensing onto a heat exchanger where cold brine passes through<br />

<strong>and</strong> recovers the latent heat (as shown in figure 3.2). The condensed vapor drips onto<br />

a distillate tray.<br />

The process is repeated all the way down the plant as both brine <strong>and</strong> distillate enter<br />

the next stage at a lower pressure. The concentrated brine is divided into two parts as<br />

it leaves the plant, the blowdown BD, which is pumped back to the sea, <strong>and</strong> a recycle<br />

stream R, which returns to the recovery section.<br />

From a mathematical point <strong>of</strong> view, the once-through design (with no reject section),<br />

<strong>and</strong> the recycle design can be represented by the same model if the zero value is set to<br />

the mass flow rates <strong>of</strong> the recycle R <strong>and</strong> the reject seawater CW streams.<br />

Furthermore, there is no distinction between heat recovery <strong>and</strong> heat rejection sections<br />

in the once-through plant.<br />

Cross-section <strong>of</strong> a stage in a typical MSF plant.<br />

Vapor<br />

Tube bundle<br />

Distillated<br />

Flashing brine<br />

Ro<strong>of</strong><br />

Demister<br />

Flash box<br />

For the recycled brine plants, the mass flow rates <strong>of</strong> the recycled brine <strong>and</strong> cooling<br />

water loops are typically 10 times greater than the distillate production rate. The latter<br />

is, in turn, approximately an order <strong>of</strong> magnitude greater than the steam supply mass<br />

flow rate.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

55


FIGURE 3.3<br />

56<br />

MSF desalination steady-state model<br />

MSF plant operation can be better analyzed by temperature pr<strong>of</strong>iles <strong>and</strong> sorting out<br />

the main parameters. The temperature pr<strong>of</strong>iles <strong>of</strong> a recycled brine plant are illustrated<br />

in figure 3.3. The first obvious parameter is the temperature range, ∆T,<br />

which is the<br />

difference between the top temperature (TB,o)<br />

<strong>of</strong> the incoming feed <strong>and</strong> cooling<br />

water, i.e. seawater, Tsea.<br />

Another important parameter is the temperature rise in the<br />

brine heater, (= TB,o<br />

– TF,1).<br />

Temperature pr<strong>of</strong>ile <strong>of</strong> a recycle brine MSF plant.<br />

Brine<br />

heater Heat recovery<br />

T S<br />

T F1<br />

T Bo<br />

Brine recirculation<br />

Flashing brine<br />

Distillate<br />

A non-uniform temperature difference is assumed over the entire flash range, but this<br />

does not imply a different design for each stage. This means that the interstage<br />

temperature differences will vary slightly down the plant <strong>and</strong> may vary significantly<br />

between stages <strong>of</strong> different designs. Specifically, the interstage temperature<br />

differences in the recovery <strong>and</strong> reject sections may differ considerably.<br />

The total temperature drop in each stage may have a number <strong>of</strong> causes, including:<br />

a) Interstage temperature difference ( δT):<br />

the drop temperature <strong>of</strong> all fluids at each<br />

stage. As a first assumption, all the fluids <strong>of</strong> an MSF plant have the same<br />

interstage temperature difference.<br />

b) Condenser terminal difference ( δTC):<br />

the temperature difference between the<br />

recycled brine flow being heated inside the evaporator tubes (being heated) <strong>and</strong><br />

the flashed vapor temperature at each stage. This value strongly depends on the<br />

heat exchanger type (design, material, fouling effect, etc.). A high heat transfer<br />

coefficient value means a lower δT<br />

value.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

C<br />

Cooling<br />

reject<br />

Heat rejection<br />

Make-up<br />

Blowdown + distillate<br />

Feedwater T sea<br />

c) Demister pressure losses ( δTP):<br />

the frictional pressure loss when the vapor is<br />

passed through the demister, to remove any entrained brine droplets, results in a<br />

further decrease in saturation temperature. The resulting saturation temperature<br />

drop can be estimated either using the Clausius-Clayperon relationship or the<br />

steam tables.


Mathematical model <strong>of</strong> MSF unit<br />

d) Condenser pressure losses: vapor undergoes a frictional pressure loss in the<br />

condenser tube bundle as when passing through the demisters.<br />

e) Boiling point elevation (BPE): Non-volatile solutes (i.e. sodium chloride)<br />

dissolved in water, raise its boiling point. The size <strong>of</strong> this raise may be predicted<br />

by considering the equilibrium between the solution <strong>and</strong> the water vapor, whose<br />

value is a function <strong>of</strong> the brine temperature <strong>and</strong> salinity. The BPE value is most<br />

<strong>of</strong>ten less than 1 ºC.<br />

f) Non equilibrium allowance (NEA): When the flashing brine stream enters a<br />

stage, it undergoes a pressure reduction. If this brine had an infinite residence<br />

time in the stage, the whole lot would cool down to the saturation temperature<br />

corresponding to the flash chamber pressure <strong>and</strong> a maximum amount <strong>of</strong> distillate<br />

would flashed <strong>of</strong>f.<br />

The energy consumption <strong>of</strong> an MSF plant is usually expressed in terms <strong>of</strong> the<br />

performance ratio PR, sometimes also called Gained Output Ratio, GOR defined<br />

previously. PR is commonly defined as kg <strong>of</strong> distillate per kg <strong>of</strong> dry saturated heating<br />

steam condensed in the brine heater without condensate subcooling. MSF plants<br />

normally have a PR value <strong>of</strong> 8 in the nominal case. The cleaning ball system is not<br />

normally installed in MSF plants but helps to avoid fouling in heat exchanger tubes,<br />

so the PR is also increased.<br />

Another measure <strong>of</strong> the energy consumption in MSF plants is sometimes expressed<br />

as the energy input to the brine heater per unit mass <strong>of</strong> distillate produced, <strong>of</strong>ten<br />

called the specific energy consumption (NC). This can be converted into a<br />

performance ratio, as defined above, by providing the steam condensing temperature<br />

in the brine heater.<br />

A large flash range as possible is desirable. Since the performance ratio improves as<br />

flash range increases, either for a fixed performance ratio (the operational efficiency<br />

increases due to a reduction in the required heat transfer surface area) or for a<br />

constant surface area. The recycled ratio is also reduced as the flash range increases,<br />

which results in a larger temperature rise in the heat input section for a fixed heat<br />

input, <strong>and</strong> a larger logarithmic mean temperature differences in the recovery section,<br />

with the corresponding reduction in the required heat transfer surface area.<br />

Seawater temperature limits the lowest temperature value in the plant. The only way<br />

to increase flash range is by raising the top temperature. This is limited by the onset<br />

<strong>of</strong> calcium sulphate scaling, <strong>and</strong> the increasing costs <strong>of</strong> additional stages.<br />

3.2 Mathematical model <strong>of</strong> MSF unit<br />

Several models <strong>of</strong> a single effect MSF plant are available (Barba, Liuzzo <strong>and</strong><br />

Tagliaferri, 1973; Darwish <strong>and</strong> Arazzini, 1989; Itahara <strong>and</strong> Stiel, 1968; Beamer <strong>and</strong><br />

Wilde, 1971; Coleman, 1971; Al Owais, Nijhawan <strong>and</strong> Budhijara, 1989; Helal,<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

57


FIGURE 3.4<br />

58<br />

MSF desalination steady-state model<br />

Medani <strong>and</strong> Soliman, 1986; Al-Mutaz, 1989; Alhumaizi, 1997; Hayakawa, Satori <strong>and</strong><br />

Konishi, 1973; Glueck <strong>and</strong> Bradshaw, 1970; Rautenbach <strong>and</strong> Buchel, 1979; Husain et<br />

al., 1993; Husain et al., 1994; Falcetta <strong>and</strong> Sciuba, 1997). In the SIMTAW model<br />

presented here, the energy <strong>and</strong> mass balances are applied to each stage <strong>of</strong> the MSF<br />

plant <strong>and</strong> guidelines <strong>and</strong> nomenclature following Helal et al. (1986), although all <strong>of</strong> it<br />

with significant modifications.<br />

Apart from assumptions considered in the next two sections, the following<br />

assumptions were introduced in the SIMTAW model:<br />

a) The product leaving any stage is salt free (distillate concentration = 0 ppm). No<br />

mist is entrained with the flashing vapor.<br />

b) No subcooling <strong>of</strong> condensate leaving the brine heater is considered. Furthermore,<br />

inlet steam to the brine heater is assumed to be saturated vapor, even though it<br />

can be slightly superheated, i.e., desuperheater model in the brine heater was not<br />

considered.<br />

c) There is no interstage model in SIMTAW. So, the effect <strong>of</strong> the flashing brine level<br />

per stage is not taken into account.<br />

Hence the mathematical equations —i.e., mass, energy <strong>and</strong> heat transfer equations—<br />

for a single stage (figure 3.4) <strong>and</strong> brine heater model (figure 3.5) are basically as<br />

follows:<br />

3.2.1 Stage model<br />

Referring to figure 3.4, the following equations can be written for stage number j at<br />

steady state.<br />

A general stage in a MSF plant.<br />

R<br />

T F,j<br />

C R<br />

Dj–1 TD,j–1 Bj–1 TB,j–1 CB,j–1 Cooling brine<br />

Distillate<br />

Flashing brine<br />

j th Stage<br />

R<br />

T F,j+1<br />

C R<br />

Dj TD,j Bj , flow rate<br />

TB,j, temperature<br />

CB,j , concentration<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Mathematical model <strong>of</strong> MSF unit<br />

Enthalpy balance on flashing brine:<br />

where B<br />

j<br />

Bj–1<br />

Hbj–1<br />

= Bj<br />

Hbj<br />

+ (Bj–1<br />

– Bj)<br />

Hvj<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(3.1)<br />

is the flashing brine flowstream in jth flash chamber (stage j), Hbj<br />

is the<br />

flashing brine enthalpy, which is a function <strong>of</strong> temperature <strong>and</strong> concentration. This<br />

property is calculated as a saturated liquid, Hvj<br />

is the saturated vapor enthalpy <strong>of</strong><br />

water in jth stage.<br />

Total material balance (water + salt):<br />

Bj–1<br />

+ Dj–1<br />

= Bj<br />

+ Dj<br />

where Dj<br />

is the distillate in the jth stage.<br />

Salt balance:<br />

Bj–1<br />

CB,j–1<br />

= Bj<br />

CB,j<br />

where CB,j<br />

is the salt concentration in the jth stage.<br />

Overall enthalpy balance:<br />

R CPR,j<br />

(TF,j<br />

– TF,j+1)<br />

= Dj<br />

CPD,j–1<br />

(TD,j–1<br />

– T*)<br />

+ Bj–1<br />

CPB,j–1<br />

(TB,j–1<br />

– T*) – Dj<br />

CPD,j<br />

(TD,j<br />

– T*)<br />

(3.2)<br />

(3.3)<br />

– Bj<br />

CPB,j<br />

(TB,j<br />

– T*) (3.4)<br />

where R is the recycled brine mass flow rate. In the Recovery Section, R depends on<br />

the required distillate <strong>and</strong> seawater temperature, but in the Reject Section the value<br />

corresponds to feed water supply (SR). CPR,j<br />

is the heat capacity <strong>of</strong> cooling brine,<br />

passing through the heat exchanger tubes, this property is a function <strong>of</strong> temperature<br />

<strong>and</strong> concentration. Although cooling brine is under high pressure, (to allow<br />

circulation inside the tubes), this property is calculated as if cooling brine were<br />

saturated liquid. CPD,j<br />

is the heat capacity <strong>of</strong> distillate, in this case, it is considered to<br />

be saturated liquid water; CPB,j<br />

is the heat capacity <strong>of</strong> flashing brine, which is<br />

assumed to be saturated liquid at flash chamber pressure in each stage. This property<br />

is calculated in a similar way to the cooling brine. T* is the temperature reference<br />

(273.15 K); TF,j<br />

is the cooling brine temperature in the jth stage; TD,j<br />

is the distillate<br />

temperature in the jth stage, <strong>and</strong> TB.j<br />

is the flashing brine temperature in the jth stage.<br />

59


60<br />

MSF desalination steady-state model<br />

Heat transfer equation (condenser):<br />

TD, j–<br />

TF, j+ 1<br />

---------------------------------<br />

TD, j–<br />

TF, j<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(3.5)<br />

where A j is the total evaporator/condenser heat exchange area; U j is the overall heat<br />

transfer coefficient <strong>of</strong> the evaporator in each stage. Its value depends on the various<br />

heat transfer resistance in the plant. The overall heat transfer coefficient is then:<br />

U j<br />

where R bi is the inside tube heat transfer resistance, given by<br />

(3.6)<br />

(3.7)<br />

where OD <strong>and</strong> ID are the outside <strong>and</strong> inside tube diameters respectively <strong>and</strong> h i is the<br />

convective heat transfer coefficient for fully-developed turbulent flow inside a tube.<br />

Assuming a small temperature difference between the wall surface <strong>and</strong> the bulk <strong>of</strong> the<br />

fluid,<br />

(3.8)<br />

where E is the ‘Enhancement factor’ (for smooth tubes this is 1.0, but is much greater<br />

for enhanced tubes); Re is the Reynolds number <strong>of</strong> the tube flow, Pr is the Pr<strong>and</strong>tl<br />

number <strong>of</strong> the tube flow.<br />

R w is the tube wall resistance, given by<br />

⎛ Uj ⋅ Aj ⎞<br />

exp ⎜---------------------- ⎟<br />

⎝R⋅CPR, j⎠<br />

where d lm is the logarithmic mean diameter <strong>of</strong> the tube, defined as:<br />

=<br />

1<br />

= ----------------------------------------------<br />

Rbi + Rw + Rc + Rf R bi<br />

=<br />

1<br />

-----hbi<br />

OD<br />

⋅ --------<br />

ID<br />

hbi E 0.023 k<br />

----- Re<br />

ID<br />

0.8 Pr 0.4<br />

= ⋅ ⋅<br />

R w<br />

d lm<br />

t ⋅ OD<br />

= -----------------kw<br />

⋅ dlm OD – ID<br />

=<br />

--------------------<br />

OD<br />

ln--------<br />

ID<br />

(3.9)<br />

(3.10)<br />

k w is the thermal conductivity <strong>of</strong> the wall <strong>and</strong> t is the wall thickness. Note that the<br />

tube wall resistance can be reduced, by either reducing the wall thickness or<br />

increasing the thermal conductivity <strong>of</strong> the wall.


Mathematical model <strong>of</strong> MSF unit<br />

R c is the resistance from the condensate film on the vapor-side, given by<br />

(3.11)<br />

where h c is the condensing film heat transfer coefficient obtained from the wellknown<br />

Nusselt equation:<br />

h c<br />

(3.12)<br />

where k is the condensate thermal conductivity; ρ is the condensate density; λ fg is the<br />

latent heat <strong>of</strong> evaporation; n represents the number <strong>of</strong> tubes in a vertical row; µ refers<br />

to the condensate viscosity; ∆T fm is the temperature difference across the film<br />

(=T s-- T w) where T s <strong>and</strong> T w are the saturated vapor <strong>and</strong> outside wall temperatures; g is<br />

the acceleration due to gravity.<br />

The condensate properties are usually evaluated at the film temperature T fm given by<br />

T fm = T s – 0.5 (T s – T w) (3.13)<br />

R f is the overall fouling resistance, which includes the inside <strong>and</strong> outside fouling<br />

resistance <strong>and</strong> the non-condensable gas resistance. It is usually provided by the heat<br />

exchanger designer <strong>and</strong> depends on the material <strong>and</strong> acid treatment applied to both<br />

sides <strong>of</strong> the tube walls <strong>and</strong> the cleaning ball system.<br />

Distillate <strong>and</strong> flashing brine temperatures correlation:<br />

TB, j<br />

R c<br />

=<br />

1<br />

---hc<br />

k<br />

= 0.729<br />

3 ρ 2 ⎛ g λfg ⎞<br />

⎜--------------------------------- ⎟<br />

⎝n µ OD ∆Tfm⎠ 0.25<br />

=<br />

TD, j+<br />

BPE + NEA + PL<br />

(3.14)<br />

where BPE is the boiling point elevation <strong>of</strong> brine with respect to pure water. As<br />

explained below, it is a function <strong>of</strong> brine temperature <strong>and</strong> concentration; NEA<br />

represents the non equilibrium allowance, which is the temperature drop due to the<br />

non infinite residence time <strong>of</strong> flashing brine in the flash chamber. PL refers to the<br />

pressure losses <strong>and</strong> includes demister <strong>and</strong> condenser pressure losses.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 61


MSF desalination steady-state model<br />

3.2.2 Brine Heater Model<br />

FIGURE 3.5 Heat input section.<br />

Brine heater performance (figure 3.5) can be described by the following equations:<br />

Saturated steam<br />

Brine heater<br />

Saturated liquid<br />

m ST<br />

T S<br />

Mass <strong>and</strong> salt balance (brine):<br />

= R , <strong>and</strong> CBo , = CR (3.15)<br />

where B o is the mass flow in the Brine Heater outlet; C B,o is the salt concentration in<br />

the Brine Heater outlet; C R is the salt concentration in recovery section.<br />

Overall enthalpy balance:<br />

Heat recovery section<br />

R<br />

T F,1<br />

C R<br />

Bo<br />

T B,o<br />

C B,o<br />

B 0<br />

Stage 1<br />

RCPH( TBo , – TF1 , ) = mST λST 62 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(3.16)<br />

where T B,o is the brine temperature in the Brine Heater outlet; CP H is the mean heat<br />

capacity <strong>of</strong> brine flowing inside the brine heater; m ST is the steam mass flow rate to<br />

the brine heater leaving the power generation plant; λ ST is the latent heat <strong>of</strong> steam to<br />

the brine heater.<br />

Heat transfer equation in the brine heater evaporator:<br />

TS – TF1 , ⎛UH⋅AH⎞ ----------------------- =<br />

exp ⎜------------------- ⎟<br />

TS – TBo , ⎝R⋅CPH⎠ (3.17)<br />

where A H is the total heat exchange area <strong>of</strong> the brine heater; U H is the overall heat<br />

transfer coefficient <strong>of</strong> the brine heater. It contains the same terms (explained in<br />

section 3.2.1), as the overall heat transfer coefficient <strong>of</strong> the evaporator in the jth stage;<br />

T S is the saturation temperature <strong>of</strong> the vapor entering to the brine heater.


Mathematical model <strong>of</strong> MSF unit<br />

3.2.3 Mixer <strong>and</strong> splitter model<br />

This model takes into account the MSF Plant configuration <strong>and</strong> the model proposed<br />

by Helal et al. (1986). In the SIMTAW model the mixing process is considered after<br />

the last stage <strong>of</strong> the reject section. As a result this last stage is considered another<br />

distillation stage with exactly the same model as the other MSF stages. For this<br />

reason, the SIMTAW model contains an explicit mixer <strong>and</strong> splitter model, completely<br />

separate from the desalination stages (see figure 3.6) which can be modeled with the<br />

equations below. Note that even though it does not exactly reflect the real physical<br />

conditions in the plant, the results are accurate enough.<br />

Mass balance (salt + water) on mixer:<br />

(3.18)<br />

where B N is the flashing brine flow in the last stage <strong>of</strong> the reject section; BD is the<br />

blowdown mass flow rate.<br />

FIGURE 3.6 Mixing <strong>and</strong> splitting points in the MSF desalination plant.<br />

R, Recycle brine<br />

Mass balance on mixer:<br />

Enthalpy balance on mixer:<br />

( BN – BD)CBN<br />

, + FCF = R CR 18 19 20 SR<br />

Seawater inlet<br />

Rejection section<br />

BD<br />

Blowdown<br />

D, Distillate<br />

F, Make-up<br />

Deareator<br />

CW<br />

Reject seawater<br />

R = F + B N – BD (3.19)<br />

R · Hb R = (B N – BD) Hb N + F · Hb DR<br />

(3.20)<br />

where Hb N, Hb R, Hb DR are respectively the enthalpy <strong>of</strong> brine leaving the reject<br />

section, recycle stream <strong>and</strong> deaerator.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 63


MSF desalination steady-state model<br />

Mass balance on reject seawater splitter:<br />

CW = SR – F (3.21)<br />

where SR is the inlet seawater into the reject section. The temper water is neglected<br />

here.<br />

3.3 Auxiliary equations<br />

Correlations <strong>of</strong> various properties used to solve the MSF SIMTAW model are<br />

included in this section. Most <strong>of</strong> thermodynamic <strong>and</strong> transport properties <strong>of</strong> pure<br />

water <strong>and</strong> steam are calculated with the same correlations used in the steam power<br />

plant model, described in Chapter 4. Correlations for calculating the brine <strong>and</strong><br />

seawater properties in the SIMTAW model are described below, but most properties<br />

can be found in technical h<strong>and</strong>books (Fabuss <strong>and</strong> Korosi, 1968; Hömig, 1978). The<br />

correlations used in the simulator are accepted here because results that they gave are<br />

reasonable when other mathematical models have been developed (Helal et al.,<br />

1986).<br />

3.3.1 Density<br />

The expression for the brine density ρ b (lb/ft 3 ) given here is valid for the range <strong>of</strong><br />

0-26% C b concentration <strong>and</strong> 40-300 ºF temperature. Pure water density was<br />

calculated (Mothershed, 1966) from the equation below with C b = 0.<br />

ρ b<br />

Another correlation can be found in Chen et al. (1973).<br />

3.3.2 Viscosity<br />

= 62.707172 + 49.364088Cb–<br />

0.43955304 10 2<br />

⋅<br />

– 0.032554667CbTb–<br />

0.46076921 10 4<br />

⋅<br />

+ 0.63240299 10 4 – ⋅<br />

Cb Tb – 2<br />

Tb – T b<br />

64 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

2<br />

(3.22)<br />

Tabulated <strong>and</strong> interpolated data (Lewis <strong>and</strong> R<strong>and</strong>al, 1961) for a given concentration<br />

C b <strong>and</strong> temperature T b are extrapolated between the range 0 < C b < 20%,<br />

0 ºC < T b < 120 ºC, to obtain brine viscosity µ b (N·s/m 2 ). Other correlations can be<br />

found in Leyendekkers (1979); Isdale, Spence <strong>and</strong> Tudhope (1971).


Auxiliary equations<br />

3.3.3 Thermal conductivity<br />

(3.23)<br />

Tabulated data (Lewis <strong>and</strong> R<strong>and</strong>al, 1961) are used, interpolating with three<br />

concentrations C b (0%, 10%, 20% weight) at different temperatures T b (up to<br />

120 ºC). As we can see in the formula, brine thermal conductivity k b (W/mK) is close<br />

to pure water conductivity (brine is about 2% less than pure water).<br />

Yusufova et al. (1978) also provides a correlation for thermal conductivity <strong>of</strong> brine.<br />

3.3.4 Heat capacity<br />

(3.24)<br />

Specific water heat capacity CP d is the equation (3.26). The correlation <strong>of</strong> brine<br />

specific heat (BTU/lb ºF) is obtained (Helal et. al, 1986) by applying a factor<br />

dependent upon the solid concentrations <strong>and</strong> temperature to the heat capacity <strong>of</strong> pure<br />

water CP d at the desired temperature (Bromley et al., 1970):<br />

where<br />

(3.25)<br />

(3.26)<br />

where T b is the brine temperature (50 ºF < T b < 200 ºF); C b the percentage <strong>of</strong> salt<br />

concentration.<br />

3.3.5 Enthalpy<br />

µ b ( 1.745 + 2.5Cb)10 3 –<br />

( 5.26 + 4Cb)10 5 – =<br />

–<br />

Tb 9 10 7 2 – 9 3<br />

8 10 Tb 3 10 11<br />

+ ⋅ – ⋅ + ⋅<br />

4<br />

– T b<br />

– T b<br />

kb 0.569118 0.00184086 Tb 7.289 10 6 – = ( +<br />

– ⋅ Tb) ( 1 – 0.2 Cb) CPb = 1.0 – Cb ( 0.011311 – 0.0000146 Tb) CPd CPd 1.0011833 6.1666652 10 5 – T 1.3999989 10 7<br />

= – ⋅ + ⋅<br />

+<br />

1.3333336 10 9 – T 3<br />

⋅<br />

For a given concentration C b, integration <strong>of</strong> the heat capacity from the reference<br />

temperature T* = 273.15 K gives the specific enthalpy (BTU/lb) <strong>of</strong> brine solution H b<br />

at T b:<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 65<br />

2<br />

– T 2


MSF desalination steady-state model<br />

where<br />

a = 1 – C b · 0.011311<br />

a 1 = a · 1.0011833<br />

3.3.6 Vapor pressure<br />

Hb a1 ( Tb – T* ) a2 ( Tb – T* ) a3 ( Tb – T* ) 3<br />

=<br />

+ +<br />

a 2<br />

a 3<br />

a 4<br />

a 5<br />

=<br />

=<br />

=<br />

=<br />

a4 ( Tb – T* ) 4<br />

+ +<br />

a5 ( Tb – T* ) 5<br />

1.1473561 10 5 –<br />

⋅ 6.1666652 10 5 –<br />

– ⋅ ⋅a<br />

----------------------------------------------------------------------------------------------<br />

2<br />

1.3999989 10 7 –<br />

⋅ 7.0669983 10 10 –<br />

– ⋅ ⋅a<br />

------------------------------------------------------------------------------------------------<br />

3<br />

1.3333336 10 9 –<br />

⋅ 1.6043987 10 12 –<br />

– ⋅ ⋅a<br />

------------------------------------------------------------------------------------------------<br />

4<br />

1.5296 10 14 –<br />

⋅<br />

---------------------------------<br />

5<br />

66 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(3.27)<br />

The following equation (Antoine correlation) describes how the vapor pressure p s <strong>of</strong><br />

saturated steam is dependant on temperature T (using the water coefficients, Reid,<br />

Prausnitz <strong>and</strong> Sherwood, 1977):<br />

ln ps = 23.196452<br />

3816.44<br />

– ----------------------<br />

T – 46.13<br />

(3.28)<br />

Equation (3.28) is used until 441 K. Above this temperature (<strong>and</strong> the critical point),<br />

the Harlacher & Braun vapor-pressure correlation is used, with the coefficients<br />

proposed by Reid et al. (1977).<br />

68695<br />

ln ps = 60.228852 – -------------- – 5.115 T + 7.875 10<br />

T<br />

3 – ps ⋅<br />

T 2<br />

ln-----<br />

(3.29)<br />

Equation (3.29) needs an iteration algorithm, for example a Newton-Raphson<br />

method. SI units must be used. No correlation is used to calculate the vapor pressure<br />

<strong>of</strong> brine solutions.


Auxiliary equations<br />

3.3.7 Boiling point elevation<br />

Data from Stoughton <strong>and</strong> Lietzke (1965) were correlated (Friedrich <strong>and</strong> Hafford,<br />

1971) to represent the boiling point rise BPE (ºF) as a function <strong>of</strong> temperature T K <strong>and</strong><br />

salt concentration C:<br />

BPE<br />

=<br />

565.757<br />

------------------ – 9.81559 + 1.54739 ln T<br />

T K<br />

k<br />

⎛337.178 ⎞<br />

– ⎜------------------– 6.41981 + 0.922753lnT T K⎟C<br />

⎝ K<br />

⎠<br />

32.681<br />

--------------- – 0.55368 0.079022 T<br />

T K C<br />

K<br />

2<br />

⎛ ⎞<br />

+ ⎜ + ln ⎟<br />

⎝ ⎠<br />

⎧ C<br />

⎫<br />

⎪----------------------------------------------------------------------------<br />

⎪<br />

⎨266919.6<br />

379.669<br />

⎪<br />

--------------------- – ------------------ + 0.334169<br />

⎬ ⋅ 1.8<br />

2<br />

⎩ T T ⎪<br />

K<br />

K<br />

⎭<br />

where T K = (T b + 460)/1.8 (K); C = (19.819 C b)/(1 – C b).<br />

(3.30)<br />

Br<strong>and</strong>oni, del Re, <strong>and</strong> Di Giacomo (1985) include correlations for BPE <strong>and</strong> other<br />

seawater properties.<br />

3.3.8 Non-equilibrium allowance<br />

Burns <strong>and</strong> Roe correlation (Omar, 1981) reported the following empirical equation<br />

for the non-equilibrium allowance (NEA), expressed as temperature loss (ºF):<br />

NEA ( 352)<br />

( Hj) 1.1 ( ∆TB, j)<br />

0.25 –<br />

ωj 10 3 –<br />

( ⋅ ) 0.5 ( TD, j)<br />

2.5 –<br />

=<br />

(3.31)<br />

where Hj is the height <strong>of</strong> brine pool in each stage (in.); ∆TBj , is the flash down per<br />

stage (TB,j–1 – TB,j), expressed in ºF; ωj the chamber load per unit width (lb·h/ft).<br />

3.3.9 Demister <strong>and</strong> other losses<br />

Omar (1981) suggests the following empirical equation to calculate the temperature<br />

loss due to the pressure drop in the demister <strong>and</strong> condenser tubes.<br />

∆TL =<br />

exp ( 1.885 – 0.0263TD, j)<br />

where ∆T L is expressed in ºF, <strong>and</strong> T D,j is the distillate temperature (ºF) in stage j.<br />

(3.32)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 67


MSF desalination steady-state model<br />

3.4 Solution algorithm<br />

MSF can be classified as a steady-state <strong>and</strong> lumped parameter model (Husain, 1999).<br />

A wide variety <strong>of</strong> iterative solution procedures for solving non-linear algebraic<br />

equations exist in the literature. In such procedures the equations are usually split into<br />

groups <strong>and</strong> then ordered by carefully choosing the iteration variables so that the large<br />

system <strong>of</strong> equations is decomposed into simpler subsystems.<br />

The methods usually applied to solve the multistage countercurrent separation<br />

problems encompassing large systems <strong>of</strong> non-linear equations are:<br />

a) Stage by stage calculations, i.e., iterative methods,<br />

b) Global methods, e.g. Newton <strong>and</strong> quasi-Newton methods,<br />

c) Linear methods (Helal et al., 1986),<br />

d) Other mathematical procedures, such as relaxation methods or a combination <strong>of</strong><br />

several methods.<br />

The procedure to simulate a MSF plant with the SIMTAW model is a global one, i.e.,<br />

the Powell hybrid method (Powell, 1964), which was also used to solve the power<br />

plant model in Chapter 4. The subroutines implemented for this method are available<br />

in internet (UTK <strong>and</strong> ORNL, 1999).<br />

FIGURE 3.7 Solution algorithm <strong>of</strong> a MSF desalination plant model.<br />

68 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Solution algorithm<br />

Figure 3.7 shows how the Powell hybrid model is applied to solve the MSF model.<br />

First, the variable array is built with the initial values included in SIMTAW, taking<br />

into account the chosen program options. Then, the Jacobian matrix is calculated<br />

using the differences <strong>of</strong> the array function, which contains the equations that perform<br />

the MSF model, included in the above sections. Finally, the variable array is updated<br />

by multiplying the Jacobian <strong>and</strong> the array function. If the values do not vary with<br />

respect to the latest iteration (that is, they are lower than the specified tolerance), the<br />

process is finished, or new updates are made until a new value <strong>of</strong> the Jacobian matrix<br />

is needed. The condition leading to a new calculation <strong>of</strong> the Jacobian matrix depends<br />

on the convergence <strong>of</strong> the iterations. Usually the Jacobian matrix is calculated when<br />

the variable array is updated five times.<br />

The criteria for convergence applied in SIMTAW has been imposed by the Powell<br />

method (Powell, 1964). The <strong>simulation</strong> is completed when the relative error between<br />

two consecutive iterations satisfies the specified tolerance:<br />

where<br />

m<br />

xj (3.33)<br />

is the calculated value <strong>of</strong> the variable j in the iteration m; is the calculated<br />

value <strong>of</strong> the variable j in the iteration m–1; x is the variable array, containing the<br />

dependent variables needed to perform the MSF plant <strong>simulation</strong>. The variable array<br />

contains the following terms:<br />

• Flashing brine temperature in each stage (T B,j).<br />

• Cooling brine temperature in each stage (T F,j).<br />

• Distillate temperature in each stage (T D,j) (it is not a variable in the inverse<br />

problem, see Section 3.6.3).<br />

• Flashing brine concentration in each stage (C B,j).<br />

• Flashing brine flow rate in each stage (B j).<br />

• Distillate flow rate in each stage (D j).<br />

• Top brine temperature (T B,o). In the TBT option this variable is not considered<br />

(see Chapter 5).<br />

• Recovery section concentration (C R).<br />

• Deaerator temperature (T DR).<br />

max ∆x ⎛ j ⎞<br />

⎜-------- m ⎟ ≤<br />

⎝ ⎠<br />

∆x j<br />

=<br />

x j<br />

10 3 –<br />

m m– 1<br />

xj – xj<br />

m– 1<br />

xj <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 69


MSF desalination steady-state model<br />

3.5 Simulation cases<br />

The MSF brine recycle flowchart (figure 3.1) has 7 (NRC + NRJ) + 13 degrees <strong>of</strong><br />

freedom as demonstrated by the number <strong>of</strong> independent equations <strong>and</strong> unknowns.<br />

The following variables are defined for an existing plant:<br />

• Number <strong>of</strong> recovery stages NRC (=17 in our case).<br />

• Number <strong>of</strong> rejection stages NRJ (3 stages).<br />

The following five variables (design data) are fixed for each stage (assuming the<br />

number <strong>and</strong> arrangement <strong>of</strong> tubes):<br />

a) heat transfer area <strong>of</strong> evaporators A j,<br />

b) tube length L j,<br />

c) stage width w j,<br />

d) outside diameter OD j; <strong>and</strong><br />

e) inside diameter ID j (or tube thickness t).<br />

The four brine heater variables (A H, L H, OD H, <strong>and</strong> ID H) are also known. The defined<br />

variables mentioned above sum up to 5 · (NRC + NRJ) + 6 specifications. Thus, if the<br />

fouling factor is also fixed in every different stage as well as the brine heater, this will<br />

result in (NRC + NRJ+1) more specifications. Furthermore, if the brine levels in the<br />

different stages are defined (NRC + NRJ variables), then the total number <strong>of</strong><br />

specifications is<br />

=<br />

5( NRC+ NRJ)<br />

+ 6 + ( NRC+ NRJ+ 1)<br />

+ ( NRC + NRJ)<br />

7( NRC+ NRJ)<br />

+ 7<br />

70 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(3.34)<br />

The above specifications limit the degrees <strong>of</strong> freedom to only 6; obtained by<br />

subtracting 7 (NRC + NRJ) + 7 from 7 (NRC + NRJ) + 13. Since the feed<br />

temperature T sea <strong>and</strong> concentration C sea will be known, only four remaining<br />

variables will have to be specified to solve the problem.<br />

Different combinations <strong>of</strong> variables can be chosen to simulate the MSF plant,<br />

depending on the objective <strong>of</strong> the <strong>simulation</strong> study. Each set (different case) has four<br />

specifications. For example, three cases are explained below:<br />

a) The first is called performance calculation. In this case the following operating<br />

variables are specified: R, CW, F, T S, T sea, C sea; distillate production, steam


Simulation cases<br />

consumption <strong>and</strong> Top Brine Temperature are solved. This case is most useful for<br />

sensitivity <strong>analysis</strong> studies, it is the <strong>simulation</strong> case implemented in SIMTAW.<br />

Two new <strong>simulation</strong> options, explained in Section 3.5.1 <strong>and</strong> 3.5.2, were also<br />

included in the SIMTAW program: TBT option <strong>and</strong> the inverse problem option,<br />

where the fouling factors were obtained by substituting in the distillate<br />

temperature pr<strong>of</strong>ile.<br />

b) In the second case, the operating parameters F, CW, T sea, C sea, T B,o <strong>and</strong> the plant<br />

capacity D N are specified; steam consumption, steam temperature <strong>and</strong> recycle<br />

brine are solved. This case may be used to investigate the possibility <strong>of</strong><br />

maintaining a specified plant capacity when the feed temperature is modified.<br />

This case it is not considered in the SIMTAW program because the recycle brine<br />

is determined by the MSF plant (design curves).<br />

c) In the third case, the parameters F, C sea, CW/R, m ST <strong>and</strong> T sea are specified. The<br />

behavior <strong>of</strong> the whole plant is analyzed when a specified amount <strong>of</strong> steam is<br />

supplied to the desalination plant by a coupled power plant. This case is not<br />

included in the SIMTAW program, taking into account the control implemented<br />

in the <strong>combined</strong> power <strong>and</strong> MSF plant.<br />

3.5.1 TBT control<br />

The MSF Plant has a TBT control (from 84 to 112 ºC), to avoid the tube scaling,<br />

which was included in the simulator option with a fixed TBT value. The rest <strong>of</strong> the<br />

variables can be affected by this option, e.g., distillate output is close to the initial<br />

value, due to the TBT/distillate correspondence (initial curves).<br />

This option reduces the number <strong>of</strong> equations. The equation governing heat transfer in<br />

the heater is rejected because the TBT is not a constraint in this equation. This is the<br />

only equation removed from the MSF plant model. As a result, a new system <strong>of</strong><br />

equations is obtained.<br />

3.5.2 Inverse problem<br />

This problem involves calculating the global heat transfer coefficient, U, <strong>and</strong> the<br />

fouling factor <strong>of</strong> all distillation stages <strong>of</strong> the MSF plant. In this <strong>simulation</strong> option the<br />

distillation temperature pr<strong>of</strong>ile is a user variable. As a consequence, the results<br />

obtained in the brine heater are less accurate than in other <strong>simulation</strong> modes.<br />

The heat transfer equations used to calculate the distillate temperature in each stage<br />

<strong>of</strong> the recovery <strong>and</strong> reject section are omitted, when solving the inverse problem<br />

because the user should provide the distillate pr<strong>of</strong>ile. The other equations included in<br />

the MSF model remain unchanged.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 71


MSF desalination steady-state model<br />

Taking into account the four possibilities in the <strong>simulation</strong> <strong>of</strong> the MSF process (TBT<br />

control, inverse problem, both or none options), there are four mathematical models<br />

implemented in SIMTAW.<br />

3.6 Initial data <strong>and</strong> <strong>simulation</strong><br />

Internal parameters <strong>of</strong> the MSF plant were calculated in the <strong>simulation</strong> model, using<br />

some design curves provided by the manufacturers (Fisia-Italimpianti, 1996):<br />

• Top Brine Temperature (TBT) as a function <strong>of</strong> seawater temperature (SWT in<br />

figure 3.8) <strong>and</strong> distillate D.<br />

• Recycle brine R as a function <strong>of</strong> seawater temperature T sea <strong>and</strong> distillate D<br />

(figure 3.9).<br />

• Feedwater (make-up F) as a function <strong>of</strong> distillate D <strong>and</strong> seawater concentration<br />

(figure 3.10).<br />

• Seawater to reject section as a function <strong>of</strong> distillate D <strong>and</strong> seawater temperature<br />

T sea (≡ SWT) (figure 3.11).<br />

FIGURE 3.8 Correspondence between the Top Brine Temperature <strong>and</strong> distillate output.<br />

Top Brine Temperature (º C)<br />

115<br />

110<br />

105<br />

100<br />

95<br />

90<br />

85<br />

80<br />

65 %<br />

SWT 32º C<br />

TBT 84 ºC<br />

SWT 28º C<br />

TBT 112 º C<br />

SWT 25º C<br />

100 % 125 %<br />

1200 1400 1600 1800 2000 2200 2400<br />

Distillate output (T/h)<br />

72 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Initial data <strong>and</strong> <strong>simulation</strong><br />

FIGURE 3.9 Brine recirculation as a function <strong>of</strong> the distillate output.<br />

Brine recirculation (T/h)<br />

20000<br />

19500<br />

19000<br />

18500<br />

18000<br />

17500<br />

17000<br />

16500<br />

16000<br />

1000 1200 1400 1600 1800 2000 2200 2400<br />

Distillate output (T/h)<br />

FIGURE 3.10 Make-up feed water as a function <strong>of</strong> the distillate output.<br />

Make-up feed (t/h)<br />

8500<br />

8000<br />

7500<br />

7000<br />

6500<br />

6000<br />

5500<br />

5000<br />

4500<br />

65 %<br />

SWT 28º C<br />

SWT 25º C<br />

SWT 32º C<br />

100 %<br />

FIGURE 3.11 Seawater to reject section as a function <strong>of</strong> the distillate output.<br />

Sea Water to Reject (T/h)<br />

18000<br />

17500<br />

17000<br />

16500<br />

16000<br />

15500<br />

15000<br />

14500<br />

14000<br />

Sea water inlet TDS: 45,000<br />

125 %<br />

1200 1400 1600 1800<br />

Distillate output (t/h)<br />

2000 2200 2400<br />

65 %<br />

SWT 25º C<br />

SWT 32º C<br />

SWT 28º C<br />

100 %<br />

125 %<br />

1000 1200 1400 1600 1800 2000 2200 2400<br />

Distillate output (T/h)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 73


MSF desalination steady-state model<br />

These curves contain the limits <strong>and</strong> the feasible operation ranges in the MSF plant.<br />

But those graphics also could be correlated by using the real data obtained from the<br />

plant managers in 1997 (WED, 1997). Figures 3.12 to 3.15 show how the correlations<br />

have been made using regression lines in a range <strong>of</strong> 2 ºC <strong>of</strong> seawater temperature.<br />

This possibility is available in SIMTAW with the option ‘Sim. with real data’.<br />

FIGURE 3.12 Top brine temperature depending on the seawater temperature <strong>and</strong> distillate production. Data<br />

collected during the year 1997.<br />

TBT (ºC)<br />

FIGURE 3.13 Recycle brine flow as a function <strong>of</strong> the seawater temperature <strong>and</strong> production. Real data collected<br />

in the MSF distillers during 1997.<br />

R (T/h)<br />

112<br />

108<br />

104<br />

100<br />

96<br />

92<br />

88<br />

20000<br />

19500<br />

19000<br />

18500<br />

18000<br />

17500<br />

TBT 26ºC<br />

TBT 28ºC<br />

TBT 30ºC<br />

TBT 32ºC<br />

TBT 34ºC<br />

TBT 36ºC<br />

1350 1550 1750 1950 2150 D (T/h) 2350<br />

R 26ºC<br />

R 28ºC<br />

R 30ºC<br />

R 32ºC<br />

R 34ºC<br />

R 36ºC<br />

1350 1550 1750 1950 2150 D (T/h) 2350<br />

Therefore, only three input parameters are needed to run the program (note that the<br />

model has only 6 degrees <strong>of</strong> freedom): distillate or Top Brine Temperature, seawater<br />

temperature <strong>and</strong> concentration (the seawater salinity concentration C sea in Arabian<br />

Gulf area is 45,000 TDS). Steam to brine heater conditions is also requested by<br />

SIMTAW, <strong>and</strong> the temper system takes into account the seawater intake temperature<br />

<strong>and</strong> flow rate.<br />

74 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Initial data <strong>and</strong> <strong>simulation</strong><br />

FIGURE 3.14 Make-up feed flow obtained for each range <strong>of</strong> seawater temperature when real data are<br />

computed. Average data <strong>of</strong> 1997.<br />

F (T/h)<br />

FIGURE 3.15 Seawater to reject flow correlations for different seawater temperatures entering the MSF plant.<br />

Data collected during the year 1997.<br />

SR (T/h)<br />

6600<br />

5600<br />

4600<br />

3600<br />

17900<br />

17700<br />

17500<br />

17300<br />

3.6.1 Fouling effect<br />

F 26ºC<br />

F 28ºC<br />

F 30ºC<br />

F 32ºC<br />

F 34ºC<br />

F 36ºC<br />

1350 1550 1750 1950 2150 D (T/h) 2350<br />

SR 26ºC<br />

SR 28ºC<br />

SR 30ºC<br />

SR 32ºC<br />

SR 34ºC<br />

SR 36ºC<br />

1350 1550 1750 1950 2150 D (T/h) 2350<br />

Design curves account for the fouling inside <strong>and</strong> outside <strong>of</strong> the tubes, without a<br />

cleaning ball system. Although the fouling values are very difficult to evaluate, they<br />

are input data in the program.<br />

The cleaning ball system can reduce the design fouling factor by five (Barthelmes <strong>and</strong><br />

Bolmer, 1996), depending on the tube material. Overall heat transfer coefficient <strong>of</strong><br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 75


MSF desalination steady-state model<br />

the evaporator is increased from ≈2,500 to ≈3,500 W/m 2 ·K, then the Performance<br />

Ratio <strong>and</strong> the steam consumption are also improved.<br />

TABLE 3.1 Fouling factors <strong>of</strong> the heat sections in MSF Plants.<br />

3.7 Summary<br />

Tube material Fouling factor (m 2 K/W)<br />

Cooper alloys 0.00005<br />

Titanium or Stainless Steels 0.00003<br />

Without On-Load Cleaning System 0.00020<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a system requires knowledge <strong>of</strong> thermodynamic states<br />

<strong>of</strong> the system under different operating conditions <strong>and</strong> circumstances <strong>of</strong> the plant.<br />

If the data acquisition system <strong>of</strong> the plant does not provide those data or the system is<br />

not an existing plant, the state <strong>of</strong> the system could be obtained by using a<br />

mathematical model describing system behavior.<br />

Energy <strong>and</strong> mass balance, <strong>and</strong> heat transfer equations compose the mathematical<br />

model <strong>of</strong> the MSF process, so it is not necessary to apply additional equations to<br />

obtain a reasonable agreement in the model results. Correlations providing<br />

thermodynamic properties <strong>of</strong> seawater are essential for accurate results. The model is<br />

solved using conventional methods <strong>and</strong> s<strong>of</strong>tware. Mathematical method differs form<br />

the original if some important parameters <strong>of</strong> the plant are introduced. Thus, the state<br />

<strong>of</strong> the plant could be achieved below different perspectives. Finally, the model has<br />

been adjusted as much as possible, in order to respond the design but also the real<br />

behavior <strong>of</strong> the MSF plant.<br />

When the thermodynamic state <strong>of</strong> the MSF plant is obtained, the state <strong>of</strong> the steam<br />

power plant is also dem<strong>and</strong>ed if the thermoeconomic <strong>analysis</strong> is going to be<br />

performed. It will be obtained by using equations described in Chapter 4.<br />

76 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


CHAPTER 4<br />

Steam power plant<br />

steady-state model<br />

In this chapter the mathematical model <strong>of</strong> the power generation system <strong>of</strong> a dualpurpose<br />

plant is described, which is implemented in the SIMTAW program (the<br />

simulator included in Chapter 5). This model can perform both a conventional energy<br />

<strong>analysis</strong> <strong>and</strong> a thermoeconomic <strong>analysis</strong> <strong>of</strong> a power plant. Thermophysical properties,<br />

such as temperature, pressure, viscosity, specific enthalpy, specific exergy, <strong>and</strong> so on,<br />

are calculated for the most significant mass <strong>and</strong> energy flow streams, together with<br />

operating parameters <strong>of</strong> different plant units, e.g., isoentropic efficiencies, heat<br />

transfer coefficients, etc. Different operating scenarios can be simulated by varying<br />

the input data <strong>and</strong> the <strong>simulation</strong> options to analyze plant behavior <strong>and</strong> the<br />

interactions among equipment.<br />

Power plants produce both electricity <strong>and</strong> process steam used in the MSF plant to<br />

produce desalted water from seawater. The co-generation concept considers the<br />

varying dem<strong>and</strong>s for power generation <strong>and</strong> process steam in the production <strong>of</strong><br />

drinking water. Continuous water production is required throughout the year, whereas<br />

the generation <strong>of</strong> electricity will be higher in summer than in winter.<br />

In the first part <strong>of</strong> this Chapter I will describe the power plant. Later, the mathematical<br />

model together with the most significant formulae <strong>and</strong> the solution algorithm <strong>of</strong> the<br />

system <strong>of</strong> equations are explained. Finally, the model solution is given in the third<br />

section. The operating modes <strong>of</strong> the co-generation plant lead to different models that<br />

are also described in the last section <strong>of</strong> this chapter.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 4.1<br />

78<br />

Steam power plant steady-state model<br />

4.1 Model description<br />

The power generation plant is a co-generation plant providing both electrical power<br />

<strong>and</strong> the steam required by the seawater desalination plant (MSF plant). The selected<br />

power plant had six turbojets, each <strong>of</strong> them at the co-generation design point<br />

3<br />

produced 122 MW <strong>of</strong> electricity <strong>and</strong> 198 MJ/s <strong>of</strong> process heat to provide 57,600 m<br />

<strong>of</strong> drinking water per day. A maximum <strong>of</strong> 6×<br />

146 MW can be delivered in generator<br />

terminals in pure condensing mode.<br />

Extraction/condensing turbines in each unit operated under constant pressure (that is,<br />

pressure at the high-pressure (HP) turbine inlet is always constant). Each <strong>of</strong> the<br />

turbines has two sections, a single flow HP section <strong>and</strong> a single flow low-pressure<br />

(LP) section. Steam extraction outlets for the seawater desalination plant <strong>and</strong><br />

extraction points for the feedwater heaters (points 3,4,5,6 <strong>and</strong> 8 in figure 4.1) are<br />

available on both turbine sections.<br />

Steam flow is an important variable determining the behavior <strong>of</strong> the power plant. If<br />

there is no steam supply for the MSF plant, the steam flows through the LP section<br />

<strong>and</strong> is returned (via a damper <strong>and</strong> bypass line).<br />

Schematic diagram <strong>of</strong> the power generation plant. Main significant flows are numbered for later<br />

descriptions <strong>and</strong> equations.<br />

Main HP steam flows from the steam generator —point 1 in figure 4.1— through the<br />

steam supply lines to the main steam emergency <strong>and</strong> control valves, which are<br />

flange-mounted onto the lower section <strong>of</strong> the HP outer casing. The steam from the<br />

valve casings to the valve chests welded onto the HP inner casing is supplied by the<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Model description<br />

lower section, <strong>and</strong> via bypass lines between the valve <strong>and</strong> turbine casing in the upper<br />

section.<br />

Afterward, the steam enters the valve chests which house the nozzle segments. It then<br />

flows via the control wheel <strong>of</strong> the HP rotor into the impulse chamber <strong>of</strong> the turbine<br />

casing. The steam exp<strong>and</strong>s through the reaction blading <strong>and</strong> enters the exhaust steam<br />

chamber <strong>of</strong> the HP section. The steam required for the seawater desalination plant is<br />

extracted via the extraction outlets in the lower exhaust section —point 6 in<br />

figure 4.1—.<br />

A certain percentage <strong>of</strong> the steam flows through the exhaust nozzles in the upper<br />

exhaust section <strong>and</strong> then through the downstream damper <strong>and</strong> bypass line to the LP<br />

section. It then flows into the LP reaction blading via the steam inlet nozzles <strong>and</strong>,<br />

after expansion, enters the condenser at the exhaust nozzles.<br />

The simple design <strong>of</strong> the high-pressure casing is based on a single shell construction<br />

with perfect rotational symmetry. All the components <strong>of</strong> the HP section are secured<br />

so that concentric alignment <strong>and</strong> unrestricted movement is maintained under all<br />

operating conditions.<br />

First <strong>and</strong> second HP turbine extractions —points 3 <strong>and</strong> 4 in figure 4.1— are fed to the<br />

HP heaters. The first HP extraction goes to the vacuum system <strong>of</strong> the MSF plant, <strong>and</strong><br />

is condensed in the condenser. The third HP extraction —point no. 5— feeds the<br />

deaerator; <strong>and</strong> finally a smaller quantity <strong>of</strong> the lowest extraction is sent to the first LP<br />

heater (the main part is sent to the desalination plant).<br />

The LP section is a st<strong>and</strong>ard single-flow design with an upstream inlet section.<br />

Depending on the operating mode <strong>of</strong> the turbojet, the steam is directed to the first<br />

blade carrier via a vertically mounted inlet steam nozzle —point no. 7— <strong>and</strong> led to<br />

the second blade carrier via a bypass, when the amount <strong>of</strong> steam to the LP turbine is<br />

large enough. The automatically controlled water injection system in the upper<br />

section <strong>of</strong> the casing provides the cooling required in specific operating modes. A<br />

rupture disc is fitted in the outer casing as a safeguard against over pressure. LP<br />

extraction —point no. 8— feeds the second LP heater.<br />

The Power Generation Plant also contains a live steam reduction pressure station, to<br />

extract the steam flow to desalination in case <strong>of</strong> turbine system failure. As seen in<br />

figure 4.1, E1 <strong>and</strong> E2 are the live steam extractions to the two connected desalination<br />

units. The reduction pressure station mixes the live steam with water feed from the<br />

feed pump (S1 to S4 in figure 4.1), to reach the optimum pressure for the MSF plant.<br />

When the turbine does not work, a new extraction E3 is needed to feed the vacuum<br />

system <strong>of</strong> the MSF units, <strong>and</strong> a fourth one, called E4 in figure 4.1, feeds the deaerator,<br />

where it is mixed with the condensate returned to the MSF units. In this way, the<br />

steam cycle is closed, via the HP feed flow to the boiler.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

79


FIGURE 4.2<br />

80<br />

Steam power plant steady-state model<br />

4.2 Mathematical model<br />

4.2.1 Steam turbines<br />

Simulation <strong>of</strong> admission properties (Salisbury, 1974) is based on the determination <strong>of</strong><br />

the mass flow coefficient, which was defined according to the Cooke’s model<br />

(Cotton, 1993; Spencer, Cotton <strong>and</strong> Cannon, 1974) <strong>and</strong> the Stodola’s Ellipse model<br />

(Stodola, 1927; Cooke, 1985). The mass flow coefficient φ is defined as:<br />

m<br />

m<br />

φ = ------- or φ = -------<br />

(4.1)<br />

p<br />

------p<br />

--<br />

T<br />

v<br />

where m is the mass flow rate (kg/s), p is the pressure (bar), T is the temperature (K)<br />

3<br />

<strong>and</strong> v is the specific volume (m /kg). The mass flow coefficient under operating<br />

conditions can be calculated as a function <strong>of</strong> the design parameters (subscript d).<br />

Thus, the admission values can be solved:<br />

2<br />

m md 1– rpd φ --------------p p<br />

------- d<br />

--------- 1 rp<br />

T<br />

2<br />

= = --------------------<br />

–<br />

T d<br />

where rp = ---- is the pressure ratio at each turbine section (see figure 4.2):<br />

Schematic diagram <strong>of</strong> a turbine section.<br />

p i<br />

m i<br />

T i<br />

p 0<br />

p i<br />

p 0<br />

T 0<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(4.2)<br />

The admission properties <strong>of</strong> the steam turbine are evaluated using a model in which<br />

the mass flow coefficient is a function <strong>of</strong> the pressure ratio in each turbine section,<br />

<strong>and</strong> is a characteristic value for each type <strong>of</strong> turbine. This model cannot be applied to<br />

the first section, due to the fixed pressure mode which controls the steam turbine<br />

operation. Therefore,


FIGURE 4.3<br />

Mathematical model<br />

φ K 1 rp 2<br />

= ⋅ –<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(4.3)<br />

Values <strong>of</strong> the constant K are obtained using the turbine admission properties for the<br />

design conditions supplied by the manufacturer. For example, the value <strong>of</strong> K4<br />

for the<br />

4th<br />

section <strong>of</strong> the high-pressure turbine can be obtained as follows:<br />

φ 4d<br />

m 4d<br />

= ------------ = K<br />

p 4 ⋅ 1– rp4d 4d<br />

------------<br />

T 4d<br />

(4.4)<br />

rd<br />

where the subscript ‘4d’ refers to the steam properties at the 3 extraction <strong>of</strong> the high<br />

pressure turbine, taken from a performance data case (ABB, 1996b).<br />

The efficiency model is also based on the mass flow coefficient. A correlation was<br />

proposed to obtain the isoentropic efficiency <strong>of</strong> a turbine section as a function <strong>of</strong> this<br />

coefficient. The design mass flow coefficients were used to solve this correlation for<br />

the different operation loads <strong>of</strong> the plant. Polynomial formulae were obtained for<br />

nd<br />

each section <strong>of</strong> the turbine. The formula corresponding to the 2 section <strong>of</strong> the highpressure<br />

turbine is a linear function <strong>of</strong> the mass flow coefficient φ2d,<br />

obtained from<br />

nd<br />

the pressure, temperature <strong>and</strong> flow <strong>of</strong> the performance data cases in the 2 section <strong>of</strong><br />

the high-pressure turbine:<br />

η 2<br />

= f( φ2d) = 0.013985 ⋅ φ2d + 0.1002<br />

Isoentropic <strong>and</strong> real expansion <strong>of</strong> the steam in a turbine section.<br />

h<br />

h 1<br />

h 2s<br />

p 1<br />

h 2<br />

p 2<br />

s<br />

2<br />

(4.5)<br />

Finally, the thermodynamic properties <strong>of</strong> the steam at each turbine section are<br />

calculated as follows (see figure 4.3):<br />

η i<br />

hi – hi+ 1<br />

=<br />

-----------------------hi<br />

– hi+ l, s<br />

(4.6)<br />

81


FIGURE 4.4<br />

82<br />

Steam power plant steady-state model<br />

where hi<br />

is the enthalpy <strong>of</strong> the inlet section, hi+1<br />

is the enthalpy <strong>of</strong> the outlet section,<br />

<strong>and</strong> hi+1,s<br />

is the enthalpy <strong>of</strong> the outlet section in an isoentropic process.<br />

The steam pressure in the lowest section <strong>of</strong> the high-pressure turbine was a fixed<br />

value, due to the pressure control applied to the desalination units. Hence, HP <strong>and</strong> LP<br />

turbines can be considered two different pieces <strong>of</strong> equipment.<br />

4.2.2 HP heat exchangers<br />

HP heat exchangers have desuperheating, condensation, <strong>and</strong> subcooling sections<br />

(ABB, 1996c). Thus, feed water is heated by exchanging the maximum quantity <strong>of</strong><br />

heat with the steam bled from the HP extractions.<br />

The model <strong>of</strong> the HP heat exchangers is based on a correlation <strong>of</strong> the terminal<br />

temperature differences (TTD) for the different existing loads, see figure 4.4. The<br />

overall heat balances are used to calculate the amount <strong>of</strong> extracted steam from the HP<br />

turbine. The overall heat transfer coefficient in each section cannot be used because<br />

<strong>of</strong> the lack <strong>of</strong> design data, except for the heat transfer coefficient in the condensing<br />

zone, which is a design data that varies with the requested load. Moreover, it is<br />

assumed that the condensate is a saturated liquid, even though some sub-cooling may<br />

occur.<br />

TTD differences in an HP heater.<br />

T<br />

TTD o<br />

Condensation section<br />

TTDi<br />

Desuperheating section Subcooling section<br />

Numerical correlations proposed by Erbes (Erbes <strong>and</strong> Gay, 1989) were used to solve<br />

the terminal temperature differences TTD (inlet/outlet) in HP heaters:<br />

∆Ti ⎛ m ⎞<br />

--------- ⎜------ ⎟<br />

∆Td ⎝md⎠ x<br />

⎛ T ⎞<br />

⎜----- ⎟<br />

⎝Td⎠ y<br />

⎛ p ⎞<br />

⎜---- ⎟<br />

⎝pd⎠ z<br />

=<br />

⋅ ⋅<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

L<br />

(4.7)


TABLE 4.1<br />

Mathematical model<br />

where ∆Ti<br />

is the inlet TTD in an HP exchanger (usually called Drain Cooling<br />

Advantage (DCA)), <strong>and</strong> m, T <strong>and</strong> P are the feedwater properties at the HP inlet. The d<br />

subscript refers to the design conditions. The x, y <strong>and</strong> z exponents were obtained<br />

from the heat balances for different loads supplied by the manufacturer (ABB,<br />

1996b). Typical x, y <strong>and</strong> z values are shown in table 4.1:<br />

Typical x, y <strong>and</strong> z coefficient values for the inlet TTD’s in an HP heater.<br />

The outlet TTD ( ∆To)<br />

(or simply called TTD) correlation contains more factors.<br />

Thus, the mass flow rate <strong>and</strong> steam pressure <strong>of</strong> the turbine extraction are also needed,<br />

in order to model the correct behavior in all cases. Typical values for the five<br />

coefficients needed in a HP heater are shown in table 4.2.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(4.8)<br />

The Erbes <strong>and</strong> Gay model (Erbes <strong>and</strong> Gay, 1989) also provides the pressure losses in<br />

the feed waterside <strong>of</strong> the HP heat exchangers:<br />

4.2.3 LP heat exchangers<br />

x y z<br />

0.64 –0.29 0.52<br />

∆T 0<br />

---------<br />

∆T d<br />

⎛ m ⎞<br />

⎜------ ⎟<br />

⎝md⎠ x<br />

⎛ T ⎞<br />

⎜----- ⎟<br />

⎝Td⎠ y<br />

⎛ p ⎞<br />

⎜---- ⎟<br />

⎝pd⎠ z<br />

⎛ mex ⎞<br />

⎜------------- ⎟<br />

⎝mex, d⎠<br />

a<br />

⎛ pex ⎞<br />

⎜----------- ⎟<br />

⎝pex, d⎠<br />

d<br />

= ⋅ ⋅ ⋅ ⋅<br />

TABLE 4.2 Typical x, y, z, a <strong>and</strong> b coefficient values for the outlet TTD’s in an HP heater.<br />

x y z A b<br />

–2.395 4.407 –0.713 0.584 0<br />

∆p<br />

--------<br />

∆pd ⎛ m<br />

------ ⎞<br />

⎝m⎠ d<br />

1.8 ⎛ T ⎞ ⎛ p ⎞<br />

⎜----- ⎟ ⎜---- ⎟<br />

⎝Td⎠ ⎝pd⎠ 1 –<br />

=<br />

(4.9)<br />

LP heat exchangers in the Steam Power Plant only have a condensation <strong>and</strong> a<br />

subcooling section (ABB, 1996c). Usually the steam flow is saturated vapor or<br />

contains a humidity fraction. Thus the feedwater is heated by extracting the maximum<br />

quantity <strong>of</strong> steam heat from the LP extraction <strong>and</strong> the lowest HP extraction.<br />

83


Steam power plant steady-state model<br />

Correlations similar to those used in the HP heaters to calculate the TTD’s (see<br />

figure 4.5) <strong>and</strong> the pressure losses were also used to model the LP heater behavior.<br />

The exponent values were also obtained from the heat balances for different loads<br />

supplied by the manufacturer (ABB, 1996b), they are shown in tables 4.3 <strong>and</strong> 4.4.<br />

FIGURE 4.5 TTD differences in an LP heater.<br />

T<br />

TTD o<br />

4.2.4 Deaerator<br />

Condensation<br />

section<br />

Subcooling<br />

section<br />

TABLE 4.3 Typical x, y <strong>and</strong> z coefficient values for the inlet TTD’s in an LP heater.<br />

x y z<br />

0.43 –0.02 0.10<br />

TABLE 4.4 Typical x, y, z, a <strong>and</strong> b coefficient values for the outlet TTD’s in a LP heater.<br />

x y z z b<br />

–0.04 18.97 –0.12 1.11 4.33<br />

A whole plant energy balance is included when modeling the deaerator <strong>and</strong> feedwater<br />

tank behavior (ABB, 1996c). Feedwater from the LP heaters, condensate from the<br />

desalination units <strong>and</strong> cooled drain from the HP heaters enter the feedwater tank, but<br />

the operating pressure is controlled by the 3 rd HP extraction.<br />

The mass flow leaving the extraction must be correlated to assure some saturated<br />

liquid is entering the feed pump. Several parameters were included to calculate the<br />

3 rd HP extraction mass flow rate, to cover the operating range designed by the<br />

manufacturer (ABB, 1996b). The proposed correlation is:<br />

84 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

TTD i<br />

L


Mathematical model<br />

mex, 3<br />

-------------mex,<br />

3d<br />

(4.10)<br />

where m 1 is the live steam mass flow rate generated in the boiler; T 5 <strong>and</strong> P 5 are the<br />

admission properties leaving the 3 rd section, m 1,des is the difference between m 1 <strong>and</strong><br />

desalination mass flow rate m des, m 1,LS is the difference between m 1 <strong>and</strong> Live Steam<br />

extraction sent to the reducing pressure station m LS, <strong>and</strong> m 1,LS,des m 1 minus<br />

desalination <strong>and</strong> live steam (to the reduction pressure station). The last three variables<br />

have a strong influence on the rest <strong>of</strong> the plant process units, which is why they were<br />

included in the above-proposed correlation. Table 4.5 shows the coefficients<br />

calculated in the last correlation.<br />

TABLE 4.5 x, y, z, a, b <strong>and</strong> c coefficient values in deaerator.<br />

4.2.5 Condenser<br />

4.2.6 Boiler<br />

⎛ m1 ⎞<br />

⎜-------- ⎟<br />

⎝m1d⎠ x ⎛ T5 ⎞<br />

⎜------- ⎟<br />

⎝T5d⎠ y<br />

⎛ p5 ⎞<br />

⎜------- ⎟<br />

⎝p5d⎠ z ⎛ mdes ⎞<br />

⎜--------------- ⎟<br />

⎝mdes, d⎠<br />

a ⎛ mLS ⎞<br />

⎜-------------- ⎟<br />

⎝mLS, d⎠<br />

b ⎛ m1LSdes , , ⎞<br />

⎜--------------------------- ⎟<br />

⎝m1LSdesd , , , ⎠<br />

c<br />

= ⋅ ⋅ ⋅ ⋅ ⋅<br />

x y z a b c<br />

Deaerator 0.121 1.091 1.905 0.206 2.588 –0.211<br />

A global energy balance was applied to develop the condenser model. Three streams<br />

enter the vapor side <strong>of</strong> the condenser: (i) exhaust steam from the low-pressure<br />

turbine, (ii) condensate from the MSF vacuum system <strong>and</strong> (iii) discharge from the<br />

ejectors. The maximum cooling seawater flow rate is at the seawater temperature.<br />

The condensate presumably discharges at the saturation temperature (ABB, 1996d).<br />

A model was used including the heat balance <strong>of</strong> the waterside <strong>of</strong> the boiler to<br />

simulate performance <strong>of</strong> the boiler (figure 4.1). The energy needed to heat the<br />

feedwater leaving the high-pressure heater No. 1 to the fixed conditions <strong>of</strong> the steam<br />

leaving the boiler was used to calculate the natural gas consumption <strong>of</strong> the boiler<br />

(LHV <strong>of</strong> natural gas is 8026 kcal/Nm 3 ). Boiler efficiency was introduced using the<br />

design data provided by the contractors (ABB, 1996a) for different operating<br />

conditions. Pressure losses on the waterside <strong>of</strong> the boiler were computed using the<br />

following equation:<br />

∆P<br />

---------<br />

∆Pd ⎛ m ⎞<br />

⎜------ ⎟<br />

⎝md⎠ 0.463<br />

⎛ T ⎞<br />

⎜----- ⎟<br />

⎝Td⎠ 0.436 –<br />

⎛pd⎞ ⎜---- ⎟<br />

⎝ p ⎠<br />

3.917 –<br />

=<br />

(4.11)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 85


Steam power plant steady-state model<br />

A more detailed model could calculate the intermediate properties inside the boiler<br />

(the boiler in study has two economizers <strong>and</strong> three superheaters, <strong>and</strong> a non reheat<br />

process inside the boiler). A detailed boiler model clearly surpasses the scope <strong>of</strong> this<br />

Ph. D. Thesis <strong>and</strong> is not necessary to perform a thermoeconomic <strong>analysis</strong> <strong>of</strong> a whole<br />

plant.<br />

4.2.7 Valves<br />

Pressure losses in valves were calculated using the BBC Thermal kit correlations<br />

(BBC, 1979):<br />

86 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(4.12)<br />

where p is the pressure <strong>of</strong> the flow entering the valve; Z is the pressure drop<br />

coefficient (constant value); DV, is the main stop valve seat diameter (m) Sa = ,<br />

the sonic area (m 2 ), with v, specific volume (m3 m<br />

/kg), α, sonic velocity (m/s), <strong>and</strong> m<br />

the mass flow inside the inside the valve.<br />

v<br />

--<br />

α<br />

4.2.7.1 Turbine control valves<br />

The inlet <strong>of</strong> the HP turbine has four control valves with some pressure losses (about<br />

4-5 bars). The main steam mass flow in the boiler is equally divided into four parts,<br />

each flowing through one <strong>of</strong> the valves. The pressure drop coefficient value (Z) was<br />

taken to be 0.38.<br />

4.2.7.2 Boiler outlet stop valve<br />

The security valve fixed at the boiler outlet had a pressure drop coefficient Z <strong>of</strong> 2.31.<br />

4.2.7.3 Boiler inlet control valve<br />

4.2.8 Pipes<br />

∆p<br />

------ = Z<br />

p<br />

⎛ Sa ⎞<br />

⎜---------- ⎟<br />

⎝ ⎠<br />

2<br />

DV 2<br />

This valve, used to control the pressure entering the boiler, had a pressure drop<br />

coefficient Z <strong>of</strong> 1.30.<br />

Significant pressure losses occur in the pipelines, e.g., pipes to the deaerator,<br />

extraction pipes or pipes to the boiler. These are calculated by applying the<br />

correlation proposed by Erbes <strong>and</strong> Gay (1989):


Mathematical model<br />

(4.13)<br />

The value <strong>of</strong> the a coefficient depends on the type <strong>of</strong> pipe <strong>and</strong> operating conditions.<br />

Table 4.6 lists the values <strong>of</strong> the applied a coefficient.<br />

TABLE 4.6 Values <strong>of</strong> the a coefficient for each pipe <strong>of</strong> the power model.<br />

1 st HP extraction<br />

4.2.9 Pumps<br />

∆p<br />

--------<br />

∆pd 2 nd HP extraction<br />

3 rd HP extraction (to deaerator)<br />

4 th HP extraction<br />

Pipe description a<br />

The pump model is based on the efficiency versus mass flow rate curves provided by<br />

the power plant manufacturers (ABB, 1996f). Energy consumption is derived from<br />

the energy balance applied to the pump, when the conditions <strong>of</strong> the water entering<br />

<strong>and</strong> leaving the pump are known.<br />

The thermodynamic properties <strong>of</strong> the water at the inlet/outlet <strong>of</strong> the feedwater <strong>and</strong><br />

condenser pump can be calculated using the isoentropic efficiency (see figure 4.6):<br />

1.95<br />

1.95<br />

1.95<br />

LP extraction 1.95<br />

Waterside <strong>of</strong> LPH No. 2 1.5<br />

Waterside <strong>of</strong> LPH No. 1 1.5<br />

LPH2 to deaerator 1.5<br />

Feed pump to HPH No. 2 1.8<br />

Waterside <strong>of</strong> LPH No. 2 1.8<br />

Waterside <strong>of</strong> LPH No. 1 1.8<br />

LPH No. 1 to Boiler 1.85<br />

1 st HP extraction<br />

η i<br />

⎛ m<br />

------ ⎞<br />

⎝m⎠ d<br />

a ⎛ T ⎞ ⎛ p ⎞<br />

⎜----- ⎟ ⎜---- ⎟<br />

⎝Td⎠ ⎝pd⎠ 1 –<br />

= ⋅ ⋅<br />

– hi =<br />

-------------------------<br />

–<br />

hi+ 1, s<br />

hi+ 1<br />

h i<br />

(4.14)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 87<br />

1.8<br />

1.95


Steam power plant steady-state model<br />

where h i is the enthalpy <strong>of</strong> the inlet water, h i+1 is the enthalpy <strong>of</strong> the outlet water, <strong>and</strong><br />

h i+1,s is the outlet water enthalpy in an isoentropic pumping process.<br />

FIGURE 4.6 Isoentropic <strong>and</strong> real compression process in a pump.<br />

h<br />

4.2.10 Gl<strong>and</strong> <strong>and</strong> seal steam system<br />

All steam flow leakages are considered <strong>and</strong> accounted for in the heat balance<br />

calculations. Gl<strong>and</strong> steam system <strong>of</strong> the power plant is described in figure 4.7.<br />

FIGURE 4.7 Gl<strong>and</strong> <strong>and</strong> seal steam system.<br />

h 1<br />

h 2s<br />

Live steam<br />

h 2<br />

p 2<br />

Martin’s formula (Martin, 1919) for steam leakage through labyrinth seals was used<br />

to calculate the leakage flows for representative designs with normal running<br />

clearances (figure 4.8):<br />

88 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

p 1<br />

HP LP<br />

m1 + m2 = Kd m 2<br />

=<br />

K′ d<br />

s<br />

( pt – pl)pt ---------------------------<br />

273.15 + Tt ( p1 – p2)p1 ----------------------------<br />

273.15 + T1 Ejector<br />

(4.15)<br />

(4.16)


Mathematical model<br />

where Kd <strong>and</strong> K′ d are constants, obtained from the design data (ABB, 1996b), <strong>and</strong><br />

1, 2 <strong>and</strong> t subscripts refer to the first <strong>and</strong> second seal in a leakage <strong>and</strong> the steam<br />

conditions inside the turbine.<br />

FIGURE 4.8 Leakage flows <strong>and</strong> seals <strong>of</strong> a steam turbine.<br />

m 2<br />

p 2<br />

The valve connecting the high <strong>and</strong> low pressure lines <strong>of</strong> the gl<strong>and</strong> steam system (see<br />

figure 4.7) is only opened in the condensing operation mode, i.e. when the turbine is<br />

working without desalination flow <strong>and</strong> only producing electricity, due to the high<br />

amount <strong>of</strong> steam lost in the HP leakage.<br />

Finally, the energy balances in the high <strong>and</strong> low pressure lines <strong>of</strong> the gl<strong>and</strong> steam<br />

system are used, to evaluate the properties <strong>of</strong> the steam flowing to the ejector.<br />

Table 4.7 shows the Kd <strong>and</strong> K′ d<br />

values obtained for the four parts <strong>of</strong> the turbine<br />

interacting with the gl<strong>and</strong> <strong>and</strong> seal steam system.<br />

TABLE 4.7 K d <strong>and</strong> K d’ constants <strong>of</strong> the gl<strong>and</strong> <strong>and</strong> seal steam system.<br />

HP Turbine. Inlet 0.02 1.392<br />

HP Turbine. Outlet 0.83 1.448<br />

LP Turbine. Inlet 1.4288 2.962<br />

LP Turbine. Outlet 1.4288 2.962<br />

4.2.11 Generator<br />

m 1<br />

shaft<br />

p 1<br />

K d<br />

turbine<br />

Generator losses were accounted for in the model to more precisely calculate the<br />

plant’s output power, using manufacturer design data (ABB, 1996e). Generator<br />

efficiency is therefore included in equation (4.17) as a function <strong>of</strong> the output power<br />

in MW:<br />

η gen (%) = (0.941 + 9.701 · 10 –4 · MW + 7.071 · 10 –6 · MW 2<br />

+ 1.771 · 10 –8 · MW 3 ) · 100 (4.17)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 89<br />

p t , T t<br />

m t<br />

K d’


Steam power plant steady-state model<br />

The excitation system was also included, using plant performance data. The<br />

excitation system losses (ESL) are calculated by a formula that depends on the output<br />

power in kW:<br />

ESL = 0.00152 * kW – 5.01645 (4.18)<br />

Therefore, the simulator can calculate the electrical <strong>and</strong> net output power produced in<br />

the power plant.<br />

4.3 Auxiliary equations<br />

The thermodynamic <strong>and</strong> transport properties in a steam power plant <strong>simulation</strong><br />

involve pure water <strong>and</strong> steam.<br />

4.3.1 Thermodynamic properties<br />

The thermodynamic properties <strong>of</strong> water can be calculated by a group <strong>of</strong> functions<br />

using equations from the IFC-1967 formulae for industrial applications. Those<br />

formulae was accepted in the Sixth International Conference about Water Properties<br />

(1967). Since then, they have become the st<strong>and</strong>ard for ASME, JSME, etc. (also see<br />

ASME, 1967; JSME, 1968).<br />

Detailed numerical methods used to solve the inverse functions can be found in Pina<br />

(1979).<br />

4.3.2 Transport properties<br />

Specific heat at constant pressure was obtained by numerical integrating the enthalpy<br />

function. Formulae used to calculate the thermal conductivity <strong>and</strong> dynamic viscosity<br />

were taken from Sangers <strong>and</strong> Watson (1986) <strong>and</strong> Yata <strong>and</strong> Minamiyama (1979).<br />

Vargaftik (1978) covers the entire range <strong>of</strong> the properties, <strong>and</strong> numerical interpolation<br />

methods were used to complete them at the proper conditions.<br />

4.4 Solution algorithm<br />

The mathematical model <strong>of</strong> the power plant is also a set <strong>of</strong> non-linear algebraic<br />

equations. There are a wide variety <strong>of</strong> iterative procedures to solve this kind <strong>of</strong><br />

problem; splitting the equations into subgroups <strong>and</strong> then solving each subsystem to<br />

create an iteration loop. Our model was not portioned into subsystems.<br />

The power plant model is solved using the Powell hybrid method (Powell, 1964), also<br />

used by SIMTAW simulator to solve the MSF plant model. It is a derivation <strong>of</strong> the<br />

90 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Solution algorithm<br />

Newton method supported by an iterative technique where non-linear terms, such as<br />

variable products <strong>and</strong> properties, are set to constant values from the latest iteration.<br />

The Powell hybrid method is applied to the whole set <strong>of</strong> equations. It requires a<br />

considerable programming effort <strong>and</strong> computer storage. Despite this, a global method<br />

provides the best solution. Solving the whole system sequentially (where it is<br />

decomposed in a set <strong>of</strong> subsystems), or linearally (where some variables are<br />

considered a linear combinations <strong>of</strong> others), does not provide a better convergence <strong>of</strong><br />

the whole system <strong>of</strong> equations.<br />

The Powell hybrid method calculates the Jacobian by a forward-difference formula,<br />

<strong>and</strong> uses a relaxation technique to update the values in a new iteration, i.e. the<br />

Jacobian does not need to be calculated in each iteration. The applied solution<br />

algorithm is available in the Subroutine HYBRID, in the NETLIB mathematical<br />

libraries (UTK <strong>and</strong> ORNL, 1999). The user should provide a subroutine containing<br />

the model functions, which are, in turn, the functions needed in the subroutine<br />

HYBRID to calculate the Jacobian applying the forward-difference approximation.<br />

In the power plant, the number <strong>of</strong> equations is much higher than the system<br />

developed to solve the desalination unit: the variable array, (with the dependant<br />

variables needed for the power plant <strong>simulation</strong>) includes the following terms<br />

corresponding to the main flowstreams <strong>of</strong> the model:<br />

• Admission properties (m, p, h, T, η, K, φ) in each section <strong>of</strong> the HP <strong>and</strong> LP turbine.<br />

• Gl<strong>and</strong> <strong>and</strong> seal steam system properties (m, h, T).<br />

• HP <strong>and</strong> LP heaters properties (m ex, p, h, T).<br />

• Condenser <strong>and</strong> deaerator values (m, X, p, h, T).<br />

• Boiler parameters (m, p, h, T).<br />

• Pressure losses in pipes <strong>and</strong> heat exchangers (∆p).<br />

Live Steam properties are kept constant to take into account the plant operation<br />

strategy (sliding pressure control is avoided). The applied convergence criterion was<br />

the same as in the SIMTAW model to solve the MSF plant: the relative error <strong>of</strong> each<br />

variable included in the variable array between two consecutive iterations must be<br />

lower than the specified tolerance. Usually, this value is set to 10 -3 but it could be<br />

considerably reduced:<br />

where<br />

max ∆x ⎛ j ⎞<br />

⎜-------- ⎟ ≤<br />

m<br />

⎝ ⎠<br />

∆x j<br />

=<br />

x j<br />

10 3 –<br />

m m– 1<br />

xj – xj<br />

(4.19)<br />

(4.20)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 91


Steam power plant steady-state model<br />

m<br />

xj m– 1<br />

xj represents the calculated value <strong>of</strong> the variable j in the iteration m.<br />

is the calculated value <strong>of</strong> the variable j in the iteration m-1.<br />

The solution algorithm adopted to solve the mathematical model by using the Powell<br />

hybrid method is shown in figure 4.9.<br />

FIGURE 4.9 Algorithm to solve the power plant model using the Powell hybrid method.<br />

4.5 Operating modes <strong>and</strong> mathematical models<br />

A wide variety <strong>of</strong> operating modes are available in the power plant, depending on the<br />

amount <strong>of</strong> required steam for the MSF desalination units, (either via the live steam<br />

reduction pressure station or via the fourth extraction <strong>of</strong> the HP turbine). Moreover, if<br />

it is not necessary to produce electricity, the system live steam-deaerator-boiler can<br />

be used to obtain the required steam for one or two desalination units.<br />

The operating modes <strong>of</strong> the steam power plant are as follows:<br />

a) Extraction mode. The most common operation mode where the plant produces<br />

electricity <strong>and</strong> also supplies steam to the MSF unit.<br />

b) Parallel mode: When the power output is less than 75 MW, the live steam reduction<br />

pressure station supplies steam with enough pressure to the MSF unit.<br />

c) Condensing mode: In this case no distilled water is produced <strong>and</strong> the plant operates<br />

as a conventional steam power plant (the power output is maximum).<br />

92 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Operating modes <strong>and</strong> mathematical models<br />

d) Desalination mode: The opposite <strong>of</strong> the condensing mode. The plant only produces<br />

distilled water <strong>and</strong> the steam turbine does not work. Thus the boiler provides<br />

the required steam to the MSF units via the steam reduction pressure<br />

station.<br />

e) Twin desalination mode: Here the boiler is in full load operation <strong>and</strong> produces<br />

steam for two MSF desalination units. This mode is unusual <strong>and</strong> the steam turbine<br />

plant does not operate either.<br />

f) Twin extraction mode: Similar to the extraction mode, but the boiler also provides<br />

steam for a second MSF desalination unit using a portion <strong>of</strong> the live steam<br />

derived from the live steam reduction pressure station.<br />

Three different mathematical models were implemented to simulate all the different<br />

operating modes included in the boiler performance data (ABB, 1996a).<br />

The models included in the power plant <strong>simulation</strong> program were the following:<br />

(i) Normal Turbine Load Model (NTL MODEL): Mass flow entering the LP turbine<br />

is between 3-125 kg/s; then the Stodola’s model is applied to simulate the LP<br />

turbine. The amount <strong>of</strong> steam required via the live steam reducting pressure<br />

station is not important if the mass flow to LP turbine is more than the<br />

specified lower limit. This model is more complex, <strong>and</strong> has the maximum<br />

number <strong>of</strong> equations.<br />

(ii) Low Turbine Load Model (LTL MODEL): Mass flow entering the LP turbine is<br />

less than the lower limit imposed previously. The Stodola´s model cannot be<br />

applied to the LP turbine, there is a compressor action at high exhaust<br />

pressures <strong>and</strong> low loads, illustrated by the stream lines in figure 4.10:<br />

FIGURE 4.10 Last stage <strong>of</strong> LP turbine acting as a compressor.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 93


Steam power plant steady-state model<br />

This model determines entry conditions at the condenser. A parametric model<br />

based on the thermal balances is then used to solve the admission properties in<br />

the LP turbine. Thus, the number <strong>of</strong> equations in the LTL model is reduced<br />

when the LP values are solved differently. However, the model has a poor<br />

stability because negative mass flows could appear during the iteration process<br />

<strong>and</strong> the program must be aborted.<br />

(iii)Non Turbine Working Model (NTW MODEL): The Power Plant is only used to<br />

supply steam to the MSF desalination units, <strong>and</strong> the HP <strong>and</strong> LP turbines are<br />

<strong>of</strong>f. Therefore, the power plant scheme is reduced to a very simple model,<br />

composed <strong>of</strong> the boiler, live steam reduction pressure station <strong>and</strong> deaerator.<br />

HP heaters are bypassed, <strong>and</strong> pressure losses are neglected. This final scheme<br />

is shown in figure 4.11:<br />

FIGURE 4.11 Power plant scheme in the NTW Model. Some flowstreams are renumbered with respect fig. 4.1.<br />

The third model is the simplest one used to describe the power plant as the number <strong>of</strong><br />

equations is considerably reduced.<br />

Operating conditions should be classified in one <strong>of</strong> the three <strong>simulation</strong> models<br />

outlined above (see table 4.8). Performance data cases included in the Design Data <strong>of</strong><br />

the Boiler (ABB, 1996a) are:<br />

1. MSL1 (Minimum stable load at 20% boiler MCR)<br />

Load at which the boiler is still able to operate continuously with rated steam<br />

properties, without the bypass system in operation <strong>and</strong> without extraction heat<br />

flow to desalination <strong>and</strong> pressure reduction station.<br />

94 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Operating modes <strong>and</strong> mathematical models<br />

2. MSL2 (Minimum stable load)<br />

Load corresponding to unit operation at 45 MW <strong>and</strong> with <strong>combined</strong> heat flow <strong>of</strong><br />

145 Gcal/h from parallel operation <strong>of</strong> turbine extraction <strong>and</strong> live steam reducing<br />

pressure station (118.66 <strong>and</strong> 26.34 Gcal/h respectively).<br />

3. MSL3 (Minimum stable load with two distillers)<br />

The turbine is at minimum stable load <strong>and</strong> the extraction heat flow is 145 Gcal/h<br />

plus 150 Gcal/h through HP pressure reduction station.<br />

4. MSL4 (Winter operation)<br />

The turbine is at minimum stable load with an extraction heat flow <strong>of</strong> 170 Gcal/h<br />

to desalination unit.<br />

5. PL65<br />

The turbine generator load is 65 MW <strong>and</strong> the extraction heat flow is 145 Gcal/h.<br />

6. PL85<br />

The turbine generator is at 85 MW <strong>and</strong> an extraction heat flow <strong>of</strong> 145 Gcal/h.<br />

7. PL115<br />

The turbine generator at 115 MW <strong>and</strong> an extraction heat flow <strong>of</strong> 145 Gcal/h.<br />

8. MCR (Maximum Continuous Rating)<br />

The turbine generator at rated steam parameters with a power output <strong>of</strong> 115 MW<br />

<strong>and</strong> an extraction heat flow <strong>of</strong> 170 Gcal/h.<br />

9. VWO<br />

Turbine swallowing capacity (all control valves open) with extraction heat flow<br />

<strong>of</strong> 170 Gcal/h.<br />

10. MR (Maximum Rating)<br />

The turbine generator at rated steam parameters, nominal control valve spindle<br />

position <strong>and</strong> no extraction heat flow to desalination.<br />

11. Boiler MCR<br />

Maximum continuous rating <strong>of</strong> boiler to be 10% above the requirement <strong>of</strong> unit<br />

MCR test mentioned in item 8.<br />

12. Boiler peak load (COC)<br />

Boiler peak load at least 5% more than boiler MCR. The extraction heat flow is<br />

170 Gcal/h to desalination <strong>and</strong> 50.8 Gcal/h to live steam reduction pressure<br />

station.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 95


Steam power plant steady-state model<br />

13. ODOB (One desalination <strong>and</strong> one boiler only)<br />

170 Gcal/h extracted through the HP reduction pressure station (a desalination<br />

unit), turbine is not in use.<br />

14. TDOB (Two desalination <strong>and</strong> one boiler only)<br />

340 Gcal/h extracted through the HP reduction pressure station (two desalination<br />

units), turbine is not in use.<br />

Table 4.8 shows the type <strong>of</strong> model applied to simulate each operating mode in the<br />

performance data:<br />

TABLE 4.8 Operating mode <strong>and</strong> mathematical model corresponding to the performance data cases.<br />

Performance data case Mathematical Model Operating mode<br />

4.6 Summary<br />

MSL1 LTL a<br />

a. Live steam temperature is 460 ºC.<br />

Condensing<br />

MSL2 LTL Parallel<br />

MSL3 LTL Twin Extraction<br />

MSL4 LTL Extraction<br />

PL65 NTL Extraction<br />

PL85 NTL Extraction<br />

PL115 NTL Extraction<br />

MCR NTL Extraction<br />

VWO NTL Extraction<br />

MR NTL Condensing<br />

MCR NTL Extraction<br />

COC NTL Twin Extraction<br />

ODOB NTW Desalination<br />

TDOB NTW Twin Desalination<br />

The thermodynamic states <strong>of</strong> the co-generation plant with the steam turbine plant <strong>and</strong><br />

the MSF unit are now permissible thanks to the mathematical models described in the<br />

previous <strong>and</strong> this chapter. The mathematical model <strong>of</strong> the steam turbine plant is in<br />

some cases very unstable, especially when the operating conditions provoke the<br />

deviation <strong>of</strong> the steam to the MSF unit <strong>and</strong> LP turbine is forced to work in unexpected<br />

conditions.<br />

96 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Summary<br />

The set <strong>of</strong> equations composing the mathematical model depending on the operation<br />

mode <strong>of</strong> the plant is solved with a global method in which the variables are<br />

simultaneously calculated.<br />

These two chapters contain the mathematical models introduced in the simulator,<br />

which is the tool that allows the use <strong>of</strong> the model’s results in the thermoeconomic<br />

<strong>analysis</strong> <strong>of</strong> the dual-purpose plant.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 97


CHAPTER 5<br />

Simulator<br />

Mathematical models used to simulate a dual-purpose plant are quite complex (see<br />

Chapters 3 <strong>and</strong> 4), so a solid basis is needed to solve them. Despite this, the SIMTAW<br />

program has been built in such way that only a few input data are essential to simulate<br />

the power <strong>and</strong> desalination plants in order to analyze plant performance. Hence, no<br />

highly qualified background is needed to use the program, although we only<br />

recommend its use to obtain a correct underst<strong>and</strong>ing <strong>of</strong> the results to technicians <strong>and</strong><br />

plant managers that have an in depth knowledge <strong>of</strong> the dual plant.<br />

Simulation <strong>of</strong> the thermodynamic processes in a dual-purpose plant is the first step to<br />

develop the <strong>Thermoeconomic</strong> Analysis <strong>of</strong> the Plant. Thermodynamic properties <strong>of</strong> the<br />

flowstreams in the plant are needed to apply the exergy balance, <strong>and</strong> to calculate the<br />

exergy costs <strong>of</strong> these flows. In this way, the complete <strong>analysis</strong> <strong>of</strong> the irreversibilities<br />

<strong>and</strong> malfunctions can be done, <strong>and</strong> the causes that generate these faults can be<br />

detected.<br />

SIMTAW is the thermoeconomic s<strong>of</strong>tware that can provide these results. It is the<br />

result <strong>of</strong> a complex project with several model developments <strong>of</strong> different complexity.<br />

A Visual Basic coded program is the user friendly interface. SIMTAW was built<br />

following those stages:<br />

1. To solve the mathematical models using an Equation Solver. In this case the EES<br />

program was used (Klein <strong>and</strong> Alvarado, 1999). Mathematical models were<br />

solved in blocks, then the whole model was connected. Relationships between<br />

variables, <strong>and</strong> independent blocks <strong>of</strong> equations were found, then the mathematical<br />

model was translated to a high-level programming language such as Fortran.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Simulator<br />

2. The dual-plant was simulated with a Fortran coded program (Micros<strong>of</strong>t Corporation,<br />

1997). This program has several files including the design data, steam <strong>and</strong><br />

brine properties, subroutines to initialize <strong>and</strong> calculate the variables, subroutines<br />

to solve the system <strong>of</strong> equations, <strong>and</strong> the algorithm which controls the whole<br />

program.<br />

3. The Dynamic Link Libraries, usually named ‘DLL’s’, are the interface between<br />

the Fortran <strong>and</strong> Visual Basic programs. Seven ‘DLL’s’ were built to develop four<br />

mathematical models included in the MSF Plant <strong>and</strong> three Power Plant models -<br />

all these mathematical models correspond to the operating modes explained in<br />

the previous chapters-.<br />

4. Finally, a Visual Basic coded program (Micros<strong>of</strong>t Corporation, 1997) was built<br />

to make the program more user friendly. This program is described in the following<br />

section.<br />

The first section <strong>of</strong> the chapter describes how to use the simulator when the<br />

thermoeconomic state <strong>of</strong> the MSF or the steam power plant is requested. But in<br />

Chapters 3 <strong>and</strong> 4 the accuracy <strong>of</strong> the mathematical models is not analyzed. Model<br />

validation is therefore included in this section, when the data flowsheets obtained<br />

from plant designers are compared with the results given by the simulator. In general,<br />

simulator calculates the properties <strong>of</strong> the main flowstreams <strong>of</strong> the dual-plant, the<br />

associated error in the calculations is very low.<br />

5.1 SIMTAW structure<br />

SIMTAW is the program that simulates the two processes involved in a well-known<br />

dual-purpose plant: the MSF <strong>and</strong> the Power Generation units. SIMTAW has a userfriendly<br />

interface that (through a set <strong>of</strong> more than 20 windows) allows the user to<br />

proceed by clicking the specified buttons. SIMTAW is built in Visual Basic 5.0, a new<br />

version only useful for 32 bits, <strong>and</strong> requires at least Windows’95. A user guide<br />

explaining how to manage the program has been implemented (Villalon, 1995), <strong>and</strong><br />

includes a very strict control over the input data introduction in order to avoid<br />

inconsistencies in the mathematical models.<br />

The two processes can be simulated independently <strong>and</strong> are driven by two different<br />

windows. The window that manages the MSF <strong>simulation</strong> is shown in figure 5.1,<br />

containing the MSF unit scheme, <strong>and</strong> seven text boxes <strong>and</strong> control buttons. In the text<br />

boxes, the user must introduce an allowed value for the following variables:<br />

1. Distillate mass flow rate (1,200-2,400 T/h) or Top Brine Temperature (84-112 ºC)<br />

in the MSF plant.<br />

2. Seawater to reject temperature (25-36 ºC).<br />

100 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 5.1<br />

SIMTAW structure<br />

3. Seawater concentration at the seawater inlet (40,000-50,000 TDS).<br />

4. Steam to brine heater temperature (80-150 ºC).<br />

5. Steam to brine heater pressure (0.8-3.0 bar).<br />

6. Sea water temperature (18-36 ºC)<br />

7. Seawater inlet flow (12,000-20,500 T/h).<br />

SIMTAW MSF process window.<br />

After these values are correctly introduced, the user must choose the TBT control<br />

option—clicking the corresponding box—, to fix the Top Brine Temperature value<br />

during the <strong>simulation</strong>. The inverse problem option also calculates the fouling factor in<br />

each stage. The third option, called Sim. With real data,<br />

includes a correlation with<br />

real data <strong>of</strong> the main mass flow rates <strong>of</strong> the MSF unit collected during the year 1997<br />

(WED, 1997).<br />

The window that manages the power plant (figure 5.2) contains the plant scheme <strong>and</strong><br />

four text boxes where the user introduces input variables needed to perform the power<br />

plant <strong>simulation</strong>:<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

101


Simulator<br />

1. Generator output (including generator losses, 50-147 MW).<br />

1<br />

2. Live Steam extractions to the reduction station (0-340 Gcal/h).<br />

3. Steam mass flow rate to the desalination units (0-189 kg/s).<br />

4. Condenser pressure (0.02-0.14 bar).<br />

Then, the user must choose one <strong>of</strong> the six operating modes in the dual-plant,<br />

depending on the power <strong>and</strong> steam dem<strong>and</strong>ed to the MSF unit(s), the operation<br />

modes are (see section 4.6 relating the operating <strong>and</strong> mathematical models <strong>of</strong> the<br />

process):<br />

• Extraction mode.<br />

• Parallel mode.<br />

• Condensing mode.<br />

• Desalination mode.<br />

• Twin desalination mode.<br />

• Twin extraction mode.<br />

The four input variables must be consistent with the selected operating mode, anyway<br />

the program will inform you which variable is out <strong>of</strong> the range specified for each<br />

operating mode.<br />

The <strong>simulation</strong> results <strong>of</strong> both processes are also presented in several windows, <strong>and</strong><br />

are resumed here:<br />

• Relevant parameters corresponding to the whole plant <strong>and</strong> to different<br />

components (fuel consumption, performance ratio, plant efficiency, specific<br />

consumption, steam consumption, etc).<br />

• Thermophysical properties <strong>of</strong> the mass flowstreams considered in the <strong>simulation</strong><br />

(the flowstreams are numbered in figures 5.1 <strong>and</strong> 5.2 respectively). In the power<br />

plant process the values <strong>of</strong> the gl<strong>and</strong> steam leakage system are also available (see<br />

section 4.3.10 for specifications). The properties are:<br />

–<br />

–<br />

–<br />

–<br />

–<br />

–<br />

Temperature.<br />

Pressure.<br />

Mass flow rate.<br />

Steam quality.<br />

Specific enthalpy.<br />

Specific entropy.<br />

1. Taking into account for the two extraction units (E1, E2) —see figure 4.1— with the same thermodynamic<br />

properties.<br />

102 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 5.2<br />

SIMTAW structure<br />

SIMTAW power plant window.<br />

–<br />

–<br />

–<br />

–<br />

–<br />

Specific exergy (thermal, mechanical <strong>and</strong> chemical contributions).<br />

Dynamic viscosity.<br />

Thermal conductivity.<br />

Specific heat.<br />

Density.<br />

• Some charts <strong>of</strong> different variables plotted by using a graphic server in SIMTAW:<br />

temperature pr<strong>of</strong>iles in the MSF stages, distillation per stage, expansion line <strong>of</strong><br />

the steam turbine.<br />

• The exergy costs <strong>of</strong> the main components <strong>of</strong> the power plant <strong>and</strong> water are shown<br />

in a window, if the fuel cost is introduced (in dollars per unit <strong>of</strong> energy) the<br />

exergoeconomic costs are also included.<br />

All these results can be saved in a text file than can be accessed by conventional<br />

applications (MS Office). The file also includes the input values <strong>and</strong> some interesting<br />

design values introduced in the simulator (tube characteristics <strong>and</strong> fouling factor in<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

103


Simulator<br />

distillers in the MSF plant, for example), <strong>and</strong> the exergy cost <strong>of</strong> the products <strong>of</strong> each<br />

component, following the productive structure that will be explained in Chapter 7.<br />

5.2 Model validation<br />

The simulator should predict the most important values <strong>of</strong> the main flowstreams <strong>of</strong><br />

the power <strong>and</strong> desalination plant with an accuracy that allows reproducing the<br />

operating conditions <strong>of</strong> the plant without using a data flowsheet for each situation.<br />

The accuracy <strong>of</strong> the simulator is tested with the data flowsheets provided by the plant<br />

managers, also called model validation <strong>of</strong> the simulator. Furthermore, when the data<br />

acquisition system <strong>of</strong> a plant is not enough to provide the data necessary for the<br />

diagnosis <strong>of</strong> the plant, a good simulator could substitute the acquisition system.<br />

The model validation is separately applied to the power <strong>and</strong> desalination plant, note<br />

that the way to calculate the thermodynamic properties in the design flowsheets is<br />

unknown, therefore an indeterminate error is structurally included in the comparative<br />

<strong>analysis</strong> (or model validation).<br />

Only a few values calculated in the simulator are also available in the data acquisition<br />

system <strong>of</strong> the power plant (this does not means that there are more signals than the<br />

system can measure, but that the recording system is limited by the plant managers):<br />

temperature <strong>and</strong> pressure <strong>of</strong> some turbine extractions, live steam conditions <strong>and</strong><br />

feedwater temperature in some heaters. Furthermore, the live steam properties are not<br />

maintained under operating conditions, <strong>and</strong> the data collected is every four hours.<br />

Consequently, no adjustment has been made to the simulator in order to achieve a<br />

more realistic set <strong>of</strong> values <strong>of</strong> the main flowstreams <strong>of</strong> the power plant.<br />

The data acquisition system <strong>of</strong> the MSF plant only provides a few data <strong>of</strong> the main<br />

controlling variables <strong>of</strong> the process every four hours (temperatures <strong>and</strong> flow rates<br />

entering <strong>and</strong> leaving the heater, recovery <strong>and</strong> reject section, <strong>and</strong> the internal<br />

parameters mentioned above). Therefore no comparison is included between the real<br />

data <strong>and</strong> the results obtained when the simulator operates with the ‘ Sim. with real<br />

data’<br />

option, that is, using the correlated internal parameters based on real<br />

experience.<br />

5.2.1 Power plant<br />

Most <strong>of</strong> the performance data cases are simulated <strong>and</strong> compared with the data<br />

provided by the plant contractors (ABB, 1996b). The first table <strong>of</strong> each comparative<br />

study shows (in different rows) the inputs <strong>of</strong> the <strong>simulation</strong> (output power W, steam to<br />

MSF unit Md, condenser pressure Pc <strong>and</strong> Live steam extraction LS); note that the<br />

output power is not exactly the same as that proposed by the contractors. This is<br />

104 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Model validation<br />

because the input power value inserted in the simulator window is only a first step to<br />

calculate the main steam flow to the boiler. Therefore, the output results try to find out<br />

the minimum difference in both live steam mass flow <strong>and</strong> the final output power for<br />

each performance case. The feed pump consumption is also included in this table<br />

(W FP). The third row shows the relative error observed in the input process.<br />

The second table shows in its first part the pressure p, temperature T <strong>and</strong> mass flow<br />

rate m <strong>of</strong> the main flowstreams <strong>of</strong> the power plant. The second part includes the<br />

values ( p′ , T′ <strong>and</strong> m′<br />

) obtained by the simulator. Finally, the third part introduces<br />

the relative error <strong>of</strong> each property <strong>of</strong> the flowstreams ( εp,<br />

εT,<br />

εm).<br />

Each flow is<br />

numbered according to the scheme followed in figure 5.2. The meaning <strong>of</strong> each<br />

performance data case is described in section 4.6. Only the values that are provided<br />

by the contractors have been compared in the table.<br />

Analyzing the model results, when the steam to LP turbine is not close to zero, that is,<br />

in performance data cases which represent partial or full load in extraction or twin<br />

extraction mode (MCR, MR, VWO, COC, PL115, PL85 performance data cases), the<br />

highest relative error is detected in the LP extraction (< 3% in any case), but the<br />

absolute difference between the simulator <strong>and</strong> data flowsheet is minimum.<br />

However, when the NTW mathematical model is applied, i.e. a minimum amount <strong>of</strong><br />

steam passes through the LP turbine (this situation correspond to MSL3 <strong>and</strong> MSL4<br />

cases, the last one is the most usual in winter operation in the Gulf Area, when the<br />

water dem<strong>and</strong> is always high but the energy consumption decrease to the 30% <strong>of</strong> the<br />

plant capacity), the relative error could reach to a 10% in the LP extraction <strong>and</strong> the<br />

steam derived to the condenser, although in those limit cases the absolute difference<br />

detected is very low. It is clear that the mathematical model applied when the steam<br />

to LP turbine is close to zero (NTW model) is more unstable than other<br />

mathematical models applied when some amount <strong>of</strong> steam passes through the LP<br />

turbine (LTW model).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

105


TABLE 5.1<br />

TABLE 5.2<br />

Simulator<br />

5.2.1.1 MCR case<br />

Input variables for the MCR (maximum continous rating, producing both electricity <strong>and</strong> water)<br />

case.<br />

W (MW) W FP (kW) Md (kg/s) Pc (bar) LS (Gcal/h)<br />

Design 122 2308 89.68 0.072 0<br />

Simulation 122.75 2262.4 89.68 0.072 0<br />

Rel. error (%) 0.611 –2.016 0.000 0.000 0.000<br />

Model validation for the MCR case.<br />

No. p (bar) T (ºC) m (kg/s) p'<br />

(bar) T'<br />

(ºC) m'<br />

(kg/s)<br />

εp<br />

(%)<br />

εT<br />

(%)<br />

106 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

εm<br />

(%)<br />

1 93 535 156.187 93 535 156.09 0.00 0.00 0.06<br />

3 28.46 365.4 10.839 28.39 363.5 10.8 0.25 0.52 0.36<br />

4 14.79 282.1 8.303 14.73 278.1 8.24 0.41 1.42 0.76<br />

5 7.235 203.2 10.989 7.213 198.6 10.94 0.30 2.26 0.45<br />

6 2.76 130.7 3.321 2.76 130.7 3.32 0.00 0.00 0.03<br />

8 0.482 80.4 2.278 0.482 80.4 2.21 0.00 0.00 2.99<br />

9 0.072 39.5 29.631 0.072 39.5 29.75 0.00 0.00 –0.40<br />

11 39.6 36.545 39.8 36.59 –0.51 –0.12<br />

12 41 36.545 41.2 36.59 –0.49 –0.12<br />

24 5.599 5.53 1.23<br />

14 78.2 36.545 78.2 36.59 0.00 –0.12<br />

23 84.2 3.321 84.4 3.32 –0.24 0.03<br />

15 128.2 36.545 128.3 36.59 –0.08 –0.12<br />

16 162.9 156.355 162.8 156.25 0.06 0.07<br />

18 164.8 156.187 164.7 156.09 0.06 0.06<br />

22 168.8 19.142 168.7 19.04 0.06 0.53<br />

19 194.6 156.187 194.4 156.09 0.10 0.06<br />

21 198.6 10.839 198.4 10.8 0.10 0.36<br />

20 230.1 156.187 230 156.09 0.04 0.06<br />

30 27.18 10.839 27.106 10.8 0.27 0.36<br />

31 14.12 8.303 14.08 8.24 0.28 0.76<br />

32 6.655 10.989 6.634 10.94 0.32 0.45<br />

33 2.677 3.321 2.677 3.32 0.00 0.03<br />

34 0.467 2.278 0.467 2.21 0.00 2.99


TABLE 5.3<br />

TABLE 5.4<br />

Model validation<br />

5.2.1.2 MR case<br />

Input variables for the MR (maximum rating, producing only electricity) performance case.<br />

W (MW) W FP (kW) Md (kg/s) Pc (bar) LS (Gcal/h)<br />

Design 146.693 2,274 0 0,135 0<br />

Simulation 146.73 2,331.4 0 0.135 0<br />

Rel. error (%) –0.025 –2.524 0.000 0.000 0.000<br />

Model validation for the MR case.<br />

No. p (bar) T (ºC) m (kg/s) p'<br />

(bar) T'<br />

(ºC) m'<br />

(kg/s)<br />

εp<br />

(%)<br />

εT<br />

(%)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

εm<br />

(%)<br />

1 93 535 156.187 93 535 156.2 0.00 0.00 –0.01<br />

3 30.58 374.5 9.254 30.52 374 9.22 0.20 0.13 0.37<br />

4 17.7 302.8 5.492 17.61 298.6 5.49 0.51 1.39 0.04<br />

5 11.23 248.9 6.012 11.17 243.3 5.92 0.53 2.25 1.53<br />

6 8.232 218.8 13.291 8.232 213.6 13.35 0.00 2.38 –0.44<br />

8 1.913 118.8 11.778 1.916 118.9 11.65 –0.16 –0.08 1.09<br />

9 0.135 51.9 110.043 0.135 51.8 110.26 0.00 0.19 –0.20<br />

11 52 135.429 52 135.58 0.00 –0.11<br />

12 53.6 135.429 53.5 135.58 0.19 –0.11<br />

24 25.069 25 0.28<br />

14 108 135.429 107.6 135.58 0.37 –0.11<br />

23 118 13.291 117.8 13.35 0.17 –0.44<br />

15 162.4 135.429 162.3 135.58 0.06 –0.11<br />

16 184.5 156.187 184.2 156.2 0.16 –0.01<br />

18 186.6 156.187 186.3 156.2 0.16 –0.01<br />

22 190.6 14.476 190.2 14.7 0.21 –1.55<br />

19 205.3 156.187 205.1 156.2 0.10 -0.01<br />

21 209.3 9.254 209 9.22 0.14 0.37<br />

20 235 156.187 234.8 156.2 0.09 -0.01<br />

30 29.71 9.254 29.62 9.22 0.30 0.37<br />

31 17.45 5.492 17.349 5.49 0.58 0.04<br />

32 11.11 6.012 11.038 5.92 0.65 1.53<br />

33 7.676 13.291 7.674 13.35 0.03 -0.44<br />

34 1.799 11.778 1.78 11.65 1.06 1.09<br />

107


TABLE 5.5<br />

TABLE 5.6<br />

Simulator<br />

5.2.1.3 PL115 case<br />

Input variables for the PL115 performance case (partial load with 115 MW <strong>of</strong> electricity <strong>and</strong> a<br />

heat extraction to MSF <strong>of</strong> 145 Gcal/h).<br />

W (MW) W FP (kW) Md (kg/s) Pc (bar) LS (Gcal/h)<br />

Design 122 2162 75.96 0.065 0<br />

Simulation 122.12 2063.1 75.96 0.065 0<br />

Rel. Error (%) –0.098 4.574 0.000 0.000 0.000<br />

Model validation for the PL115 performance data case.<br />

No. p (bar) T (ºC) m (kg/s) p'<br />

(bar) T'<br />

(ºC) m'<br />

(kg/s)<br />

1<br />

3<br />

4<br />

5<br />

6<br />

8<br />

9<br />

11<br />

12<br />

24<br />

14<br />

23<br />

15<br />

16<br />

18<br />

22<br />

19<br />

21<br />

20<br />

30<br />

31<br />

32<br />

33<br />

34<br />

εp<br />

(%)<br />

εT<br />

(%)<br />

108 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

εm<br />

(%)<br />

93 535 148.923 93 535 148.11 0.00 0.00 0.55<br />

26.99 360.5 10.252 26.791 358.4 10.17 0.74 0.58 0.80<br />

13.97 277.4 8.03 13.848 273.2 8.07 0.87 1.51 –0.50<br />

6.749 198 10.897 6.705 193.3 10.59 0.65 2.37 2.82<br />

2.39 126 3.317 2.39 126 3.29 0.00 0.00 0.81<br />

0.588 85.4 3.237 0.587 85.4 3.13 0.17 0.00 3.31<br />

0.065 37.5 36.083 0.065 37.7 35.75 0.00 –0.53 0.92<br />

37.6 43.949 37.7 43.48 –0.27 1.07<br />

38.8 43.949 37.7 43.48 2.84 1.07<br />

6.553 6.51 0.66<br />

82.2 43.949 82.2 43.48 0.00 1.07<br />

88.7 3.317 88.9 3.29 –0.23 0.81<br />

123.2 43.949 123.4 43.48 –0.16 1.07<br />

159.7 149.087 158.9 148.27 0.50 0.55<br />

161.6 148.923 160.6 148.11 0.62 0.55<br />

165.5 18.282 164.5 18.24 0.60 0.23<br />

191.9 148.923 191.4 148.11 0.26 0.55<br />

195.8 10.252 195.3 10.17 0.26 0.80<br />

227.3 148.923 226.9 148.11 0.18 0.55<br />

25.79 10.252 25.598 10.17 0.74 0.80<br />

13.32 8.03 13.188 8.07 0.99 –0.50<br />

6.141 10.897 6.135 10.59 0.10 2.82<br />

2.295 3.317 2.299 3.29 –0.17 0.81<br />

0.563 3.237 0.562 3.13 0.18 3.31


TABLE 5.7<br />

TABLE 5.8<br />

Model validation<br />

5.2.1.4 PL85 case<br />

Input variables for the PL85 performance case (partial load with 85 MW <strong>of</strong> electricity <strong>and</strong> 145<br />

Gcal/h <strong>of</strong> extraction heat flow).<br />

W (MW) W FP (kW) Md (kg/s) Pc (bar) LS (Gcal/h)<br />

Design 91 1,649 75.62 0.055 0<br />

Simulation 91.24 1,540.4 75.62 0.05 0<br />

Rel. error (%) –0.264 6.586 0.000 0.000 0.000<br />

Model validation for the PL85 performance case.<br />

No. p (bar) T (ºC) m (kg/s) p'<br />

(bar) T'<br />

(ºC) m' (kg/s) εp (%) εT (%) εm (%)<br />

1 93 535 117.391 93 535 117.03 0.00 0.00 0.31<br />

3 21.15 340.7 7.331 21.064 340.4 7.29 0.41 0.09 0.56<br />

4 11.1 261.7 5.719 11.036 259.2 5 .83 0.58 0.96 –1.94<br />

5 5.56 187.7 7.875 5.538 184.6 7.68 0.40 1.65 2.48<br />

6 2.39 126 2.21 2.39 126 2.18 0.00 0.00 1.36<br />

8 0.261 66 0.993 0.262 66.1 0.95 –0.38 –0.15 4.33<br />

9 0.055 34.6 16.478 0.055 34.6 16.63 0.00 0.00 –0.92<br />

11 34.6 20.993 34.7 20.76 –0.29 1.11<br />

12 37 20.993 37.1 20.76 –0.27 1.11<br />

24 3.203 3.12 2.59<br />

14 65 20.993 65 20.76 0.00 1.11<br />

23 69.7 2.21 70.1 2.18 -0.57 1.36<br />

15 124.4 20.993 124.5 20.76 -0.08 1.11<br />

16 153.2 117.539 152.5 117.18 0.46 0.31<br />

18 155 117.391 154.1 117.03 0.58 0.31<br />

22 158.5 13.051 157.6 13.12 0.57 -0.53<br />

19 182.4 117.391 182.3 117.03 0.05 0.31<br />

21 185.9 7.331 185.8 7.29 0.05 0.56<br />

20 215 117.391 215 117.03 0.00 0.31<br />

30 20.39 7.331 20.31 7.29 0.39 0.56<br />

31 10.7 5.719 10.62 5.83 0.75 –1.94<br />

32 5.186 7.875 5.185 7.68 0.02 2.48<br />

33 2.348 2.21 2.347 2.18 0.04 1.36<br />

34 0.257 0.993 0.258 0.95 –0.39 4.33<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

109


Simulator<br />

5.2.1.5 MSL2 case<br />

TABLE 5.9 MSL2 performance case (minimum stable load with 45 MW <strong>of</strong> electricity <strong>and</strong> a <strong>combined</strong> heat<br />

extraction flow <strong>of</strong> 145 Gcal/h). Main input data.<br />

W (MW) W FP (kW) Md (kg/s) Pc (bar) LS (Gcal/h)<br />

Design 51 1,305 60.4 0.048 26.34<br />

Simulation 51.57 1,250.7 60.4 0.048 26.34<br />

Rel. error (%) –1.118 4.161 0.000 0.000 0.000<br />

TABLE 5.10 Model validation for the MSL2 performance case.<br />

No. p (bar) T (ºC) m (kg/s) p' (bar) T' (ºC) m' (kg/s) εp (%) εT (%) εm (%)<br />

1 93 535 86.5 93 535 86.49 0.00 0.00 0.01<br />

3 13.75 324.9 4.477 13.714 322.9 4.49 0.26 0.62 –0.29<br />

4 7.442 251.2 3.355 7.411 247.4 3.38 0.42 1.51 –0.75<br />

5 4.074 186.8 4.971 4.061 182.6 4.89 0.32 2.25 1.63<br />

6 2.39 137.9 0.454 2.39 135.8 0.46 0.00 1.52 –1.32<br />

8 0.054 34.4 0 0.055 34.7 0 –1.85 –0.87 0.00<br />

9 0.048 80 1.751 0.048 80 1.75 0.00 0.00 0.06<br />

11 32.4 3.466 32.2 3.55 0.62 –2.42<br />

12 46.9 3.466 47.4 3.52 –1.07 –1.56<br />

24 0.454 0.46 –1.32<br />

14 47.3 3.466 47.7 3.52 –0.85 –1.56<br />

23 49.6 0.454 50.2 0.46 –1.21 –1.32<br />

15 125.7 3.466 125.7 3.52 0.00 –1.56<br />

16 142.4 90.218 142 90.19 0.28 0.03<br />

18 144.3 86.5 143.7 86.49 0.42 0.01<br />

22 147.4 7.832 146.7 7.87 0.47 –0.49<br />

19 166.4 86.5 166.1 86.49 0.18 0.01<br />

21 169.5 4.477 169.2 4.49 0.18 –0.29<br />

20 194.5 86.5 194.4 86.49 0.05 0.01<br />

30 13.32 4.477 13.288 4.49 0.24 –0.29<br />

31 7.235 3.355 7.206 3.38 0.40 –0.75<br />

32 3.871 4.971 3.864 4.89 0.18 1.63<br />

33 2.388 0.454 2.387 0.46 0.04 –1.32<br />

34 0.108 0 0.055a 0 49.07 0.00<br />

a. Note that the simulator does not suppose a pressure loss in the 5 th extraction if any vapor is extracted.<br />

110 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Model validation<br />

5.2.1.6 MSL3 case<br />

TABLE 5.11 Input data <strong>of</strong> the MSL3 performance case (minimum stable load with two extractions <strong>of</strong> 150 <strong>and</strong><br />

145 Gcal/h to MSF units).<br />

W (MW) W FP (kW) Md (kg/s) Pc (bar) LS (Gcal/h)<br />

Design 72.44 2,975 75.13 0.048 150<br />

Simulation 73.6 3,122.2 75.2 0.048 150<br />

Rel. error (%) –1.601 –4.948 –0.093 0.000 0.000<br />

TABLE 5.12 Model validation for the MSL3 performance case.<br />

No. p (bar) T (ºC) m (kg/s) p' (bar) T' (ºC) m' (kg/s) εp (%) εT (%) εm (%)<br />

1 93 535 163 93 535 163.1 0.00 0.00 –0.06<br />

3 18.22 333 9.851 18.176 329.8 9.94 0.24 0.96 –0.90<br />

4 9.344 252.4 7.883 9.314 247.4 7.69 0.32 1.98 2.45<br />

5 4.891 179.6 10.272 4.687 174.6 10.02 4.17 2.78 2.45<br />

6 2.39 126 0.461 2.39 126 0.5 0.00 0.00 –8.46<br />

8 0.054 34.3 0 0.055 34.7 0 –1.85 –1.17 0.00<br />

9 0.048 80 1.727 0.048 80 1.91 0.00 0.00 –10.60<br />

11 32.4 3.462 32.2 3.75 0.62 –8.32<br />

12 47 3.462 46.4 3.72 1.28 –7.45<br />

24 0.461 0.5 –8.46<br />

14 47.3 3.462 46.7 3.72 1.27 –7.45<br />

23 49.6 0.461 49.3 0.5 0.60 –8.46<br />

15 125.7 3.462 125.7 3.72 0.00 –7.45<br />

16 142.8 183.567 142 183.61 0.56 –0.02<br />

18 144.6 163 144.2 163.1 0.28 –0.06<br />

22 148.7 17.514 148.4 17.63 0.20 –0.66<br />

19 171.6 163 171.3 163.1 0.17 –0.06<br />

21 175.6 9.851 175.5 9.94 0.06 –0.90<br />

20 204 163 204.1 163.1 –0.05 –0.06<br />

30 16.6 9.851 16.62 9.94 –0.12 –0.90<br />

31 8.465 7.663 8.505 7.69 –0.47 –0.35<br />

32 3.907 10.272 4.026 10.02 –3.05 2.45<br />

33 2.388 0.461 2.387 0.5 0.04 –8.46<br />

34 0.108 0 0.055a 0 49.07 0.00<br />

a. The same argumentation <strong>of</strong> the MSL2 case.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 111


Simulator<br />

5.2.1.7 MSL4 case<br />

TABLE 5.13 Input data <strong>of</strong> the MSL4 performance case (minimum stable load with the maximum heat flow<br />

extraction to MSF unit: 170 Gcal/h).<br />

W (MW) W FP (kW) Md (kg/s) Pc (bar) LS (Gcal/h)<br />

Design 75.52 1,543 88.63 0.021 0<br />

Simulation 76.36 1,501.5 88.63 0.021 0<br />

Rel. error (%) –1.112 2.690 0.000 0.000 0.000<br />

TABLE 5.14 MSL4 performance case. Model validation.<br />

No. p (bar) T (ºC) m (kg/s) p' (bar) T' (ºC) m' (kg/s) εp (%) εT (%) εm (%)<br />

1 93 535 109.5 93 535 109.64 0.00 0.00 –0.13<br />

3 19.87 340.3 6.537 19.821 338.1 6.55 0.25 0.65 –0.20<br />

4 10.58 262.9 4.956 10.534 258.9 5.04 0.43 1.52 –1.69<br />

5 5.514 192.5 6.789 5.495 188.1 6.68 0.34 2.29 1.61<br />

6 2.76 130.7 0.424 2.76 131.1 0.45 0.00 –0.31 –6.13<br />

8 0.025 20.9 0 0.024 20.5 0 4.00 1.91 0.00<br />

9 0.021 80 1.004 0.021 79.8 1.14 0.00 0.25 –13.55<br />

11 18.3 2.743 18.3 2.92 0.00 –6.45<br />

12 36.7 2.743 36.6 2.92 0.27 –6.45<br />

24 0.424 0.45 –6.13<br />

14 37.1 2.736 36.9 2.9 0.54 –5.99<br />

23 39.2 0.424 39.1 0.45 0.26 –6.13<br />

15 130.5 2.736 130.5 2.9 0.00 –5.99<br />

16 153.6 109.649 153 109.79 0.39 –0.13<br />

18 155.3 109.5 154.7 109.64 0.39 –0.13<br />

22 158.8 11.493 158.2 11.58 0.38 –0.76<br />

19 180.8 109.5 180.7 109.64 0.06 –0.13<br />

21 184.2 6.537 184.1 6.55 0.05 –0.20<br />

20 212.2 109.5 212.2 109.6 0.00 –0.09<br />

30 19.23 6.537 19.176 6.55 0.28 –0.20<br />

31 10.26 4.956 10.207 5.04 0.52 –1.69<br />

32 5.233 6.789 5.22 6.68 0.25 1.61<br />

33 2.759 0.424 2.758 0.45 0.04 –6.13<br />

34 0.063 0 0.024a 0 61.90 0.00<br />

a. No pressure losses are associated to the final extraction<br />

112 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Model validation<br />

5.2.1.8 ODOB case<br />

TABLE 5.15 Main input data <strong>of</strong> the ODOB case (one desalination-one boiler).<br />

W (MW) W FP (kW) Md (kg/s) Pc (bar) LS (Gcal/h)<br />

Design 0 ? 88.45 0 170<br />

Simulation 0 1,222.6 88.45 0 170<br />

Rel. error (%) 0.000 ? 0.000 0.000 0.000<br />

TABLE 5.16 Model validation <strong>of</strong> the ODOB case.<br />

No. p (bar) T (ºC) m (kg/s) p' (bar) T' (ºC) m' (kg/s) εp (%) εT (%) εm (%)<br />

1 93 535 70.383 93 535 70.38 0.00 0.00 0.00<br />

3 0 0 0 0 0 0<br />

4 0 0 0 0 0 0<br />

5 0 0 0 0 0 0<br />

6 0 0 0 0 0 0<br />

8 0 0 0 0 0 0<br />

9 0 0 0 0 0 0<br />

11 0 0 0 0<br />

12 0 0 0 0<br />

24 0 0<br />

14 0 0 0 0<br />

23 0 0 0 0<br />

15 0 0 0 0<br />

16 138.9 93.841 138.9 94.94 0.00 –1.17<br />

18 140.7 70.383 140.5 70.38 0.14 0.00<br />

22 0 0 0 0<br />

19 140.7 70.383 140.5 70.38 0.14 0.00<br />

21 0 0 0 0<br />

20 140.7 70.383 140.5 70.38 0.14 0.00<br />

30 0 0 0 0<br />

31 0 0 0 0<br />

32 0 0 0 0<br />

33 0 0 0 0<br />

34 0 0 0 0<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 113


Simulator<br />

5.2.1.9 TDOB case<br />

TABLE 5.17 Main input data <strong>of</strong> the TDOB case (two desalination-one boiler).<br />

W (MW) W FP (kW) Md (kg/s) Pc (bar) LS (Gcal/h)<br />

Design 0 ? 140.766 0 340<br />

Simulation 0 764.8 140.76 0 340<br />

Rel. error (%) 0.000 ? 0.004 0.000 0.000<br />

TABLE 5.18 Model validation data for the TDOB case.<br />

No. p (bar) T (ºC) m (kg/s) p' (bar) T' (ºC) m' (kg/s) εp (%) εT (%) εm (%)<br />

1 93 535 140.766 93 535 140.76 0.00 0.00 0.00<br />

3 0 0 0 0 0 0<br />

4 0 0 0 0 0 0<br />

5 0 0 0 0 0 0<br />

6 0 0 0 0 0 0<br />

8 0 0 0 0 0 0<br />

9 0 0 0 0 0 0<br />

11 0 0 0 0<br />

12 0 0 0 0<br />

24 0 0<br />

14 0 0 0 0<br />

23 0 0 0 0<br />

15 0 0 0 0<br />

16 138.9 187.682 138.9 187.38 0.00 0.16<br />

18 140.8 140.766 139.7 140.76 0.78 0.00<br />

22 0 0 0 0<br />

19 140.8 140.766 139.7 140.76 0.78 0.00<br />

21 0 0 0 0<br />

20 140.8 140.766 139.7 140.76 0.78 0.00<br />

30 0 0 0 0<br />

31 0 0 0 0<br />

32 0 0 0 0<br />

33 0 0 0 0<br />

34 0 0 0 0<br />

114 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Model validation<br />

5.2.1.10 VWO case<br />

TABLE 5.19 Main input data <strong>of</strong> the VWO performance case (maximum capacity <strong>of</strong> the steam turbine with <strong>and</strong><br />

extraction heat flow <strong>of</strong> 170 Gcal/h to MSF).<br />

W (MW) W FP (kW) Md (kg/s) Pc (bar) LS (Gcal/h)<br />

Design 126.587 2,412 89.69 0.074 0<br />

Simulation 126.78 2,425.9 89.68 0.074 0<br />

Rel. error (%) –0.152 –0.576 0.011 0.000 0.000<br />

TABLE 5.20 Model validation data for the VWO case.<br />

No. p (bar) T (ºC) m (kg/s) p' (bar) T' (ºC) m' (kg/s) εp (%) εT (%) εm (%)<br />

1 93 535 161.038 93 535 160.21 0.00 0.00 0.51<br />

3 29.36 368.1 11.317 29.158 365.4 11.17 0.69 0.73 1.30<br />

4 15.23 284.3 8.665 15.11 279.6 8.59 0.79 1.65 0.87<br />

5 7.419 204.7 11.473 7.368 199.5 11.33 0.69 2.54 1.25<br />

6 2.759 130.7 3.485 2.76 130.7 3.44 –0.04 0.00 1.29<br />

8 0.533 82.9 2.623 0.528 82.7 2.4 0.94 0.24 8.50<br />

9 0.074 40 32.64 0.074 40.1 32.44 0.00 –0.25 0.61<br />

11 40.1 40.064 40.1 39.63 0.00 1.08<br />

12 41.3 40.064 42.2 39.6 –2.18 1.16<br />

24 6.109 5.85 4.24<br />

14 80.4 40.064 80.3 39.6 0.12 1.16<br />

23 86.6 3.485 86.8 3.44 –0.23 1.29<br />

15 128 40.064 128.1 39.6 –0.08 1.16<br />

16 163.8 161.209 163.5 160.38 0.18 0.51<br />

18 165.7 161.038 165.5 160.21 0.12 0.51<br />

22 169.7 19.982 169.5 19.76 0.12 1.11<br />

19 195.8 161.038 195.6 160.21 0.10 0.51<br />

21 199.9 11.317 199.6 11.17 0.15 1.30<br />

20 231.8 161.038 231.4 160.21 0.17 0.51<br />

30 28.01 11.317 27.817 11.17 0.69 1.30<br />

31 14.53 8.665 14.41 8.59 0.83 0.87<br />

32 6.8 11.473 6.758 11.33 0.62 1.25<br />

33 2.667 3.485 2.671 3.44 –0.15 1.29<br />

34 0.515 2.623 0.512 2.4 0.58 8.50<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 115


Simulator<br />

5.2.1.11 COC case<br />

TABLE 5.21 Input data <strong>of</strong> the COC performance case (boiler peak load at least 5% more than the MCR case).<br />

W (MW) W FP (kW) Md (kg/s) Pc (bar) LS (Gcal/h)<br />

Design 124.41 3,103 89.72 0.072 50.8<br />

Simulation 124.65 3,350 89.72 0.072 50.8<br />

Rel. error (%) –0.193 –7.960 0.000 0.000 0.000<br />

TABLE 5.22 Model validation data for the COC case.<br />

No. p (bar) T (ºC) m (kg/s) p' (bar) T' (ºC) m' (kg/s) εp (%) εT (%) εm (%)<br />

1 93 535 180.556 93 535 179.58 0.00 0.00 0.54<br />

3 29.02 366.9 12.704 28.77 363.7 12.6 0.86 0.87 0.82<br />

4 14.93 282.3 9.723 14.786 277 9.7 0.96 1.88 0.24<br />

5 7.219 202.2 12.96 7.17 196.8 12.55 0.68 2.67 3.16<br />

6 2.76 130.7 3.314 2.76 130.7 3.29 0.00 0.00 0.72<br />

8 0.479 80.3 2.258 0.471 79.9 2.17 1.67 0.50 3.90<br />

9 0.072 39.5 29.449 0.072 39.5 29.14 0.00 0.00 1.05<br />

11 39.6 36.335 39.6 35.92 0.00 1.14<br />

12 41 36.335 40.7 35.92 0.73 1.14<br />

24 5.572 5.46 2.01<br />

14 78.1 36.335 77.7 35.92 0.51 1.14<br />

23 84 3.314 83.9 3.29 0.12 0.72<br />

15 128.2 36.335 128.3 35.92 –0.08 1.14<br />

16 161.4 187.942 160.3 186.94 0.68 0.53<br />

18 163.5 180.556 162.8 179.58 0.43 0.54<br />

22 167.7 22.427 167 22.3 0.42 0.57<br />

19 193.8 180.556 193.3 179.58 0.26 0.54<br />

21 198 12.704 197.5 12.6 0.25 0.82<br />

20 230 180.556 229.6 179.58 0.17 0.54<br />

30 27.28 12.704 27.06 12.6 0.81 0.82<br />

31 14.02 9.723 13.888 9.7 0.94 0.24<br />

32 6.398 12.96 6.415 12.55 –0.27 3.16<br />

33 2.677 3.314 2.678 3.29 –0.04 0.72<br />

34 0.464 2.258 0.457 2.17 1.51 3.90<br />

116 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Model validation<br />

5.2.2 MSF Plant<br />

Distiller design data in the most characteristic operating conditions have been<br />

provided by the plant manufacturers (Italimpianti, 1997). Only a few cases contain<br />

the temperature pr<strong>of</strong>ile <strong>of</strong> the three flows inside each stage <strong>of</strong> the distiller. They<br />

correspond to the guarantied conditions <strong>of</strong> the contractors:<br />

• Nominal production in summer (normal-temperature operation in summer,<br />

NTOS): 1,900 T/h <strong>of</strong> freshwater produced (or a TBT <strong>of</strong> 100 ºC) <strong>and</strong> a seawater<br />

temperature <strong>of</strong> 32 ºC.<br />

• Maximum production in summer (high-temperature operation in summer,<br />

HTOS): distillation <strong>of</strong> 2,258 T/h (112 ºC TBT) with a seawater entering at 32 ºC.<br />

• Minimum production in summer (low-temperature operation in summer, LTOS):<br />

distillation <strong>of</strong> 1,232 T/h (84 ºC TBT), seawater enters at 32 ºC.<br />

• Maximum production in winter (high-temperature operation in winter, HTOW):<br />

distillation <strong>of</strong> 2,400 T/h (112 ºC TBT) with a seawater entering at 18 ºC. Seawater<br />

to reject section enters at 25 ºC by using the temper system by the way <strong>of</strong><br />

deviating a quantity <strong>of</strong> cooling seawater rejected to the sea.<br />

The first table <strong>of</strong> each comparative study shows some inputs <strong>of</strong> design data <strong>and</strong><br />

<strong>simulation</strong> in the first <strong>and</strong> second rows respectively (seawater intake flow SW <strong>and</strong><br />

temperature T sea). Some other inputs (steam to heater conditions, seawater intake<br />

temperature) needed for the simulator are not included because they must be the same<br />

quantity as the proposed design value. The distillate produced in the two cases is<br />

maintained in the same quantity too. Other operating parameters that are obtained in<br />

the <strong>simulation</strong> are also compared in the table: seawater to reject <strong>and</strong> recycle brine<br />

flows (SR <strong>and</strong> R), Top Brine Temperature (TBT), Performance Ratio (PR) <strong>and</strong> steam<br />

consumption (m ST). The third row shows the relative error observed in the table, the<br />

highest error is in the steam consumed. This error can be due to the absence <strong>of</strong> a<br />

desuperheater before the brine heater in the mathematical model applied to the MSF<br />

distillers, <strong>and</strong> the error introduced when the steam properties (the latent heat <strong>of</strong><br />

vaporization) below two different perspectives are calculated.<br />

The second table shows in its first part the chamber pressure p, temperature pr<strong>of</strong>ile<br />

(cooling brine TF, distillate TD <strong>and</strong> flashing brine TB) <strong>and</strong> distillate flow rate (D) <strong>of</strong><br />

each stage <strong>of</strong> the MSF plant. The second part includes the values ( p′ , TF′ , TD′ , TB′ <strong>and</strong> D′<br />

) obtained by the simulator. Finally, the third part introduces the relative error<br />

<strong>of</strong> the stage values (εp, εTF, εTD, εTB, εD). Each stage is numbered according to the<br />

scheme followed in figure 5.1.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 117


Simulator<br />

In Gulf Area the water dem<strong>and</strong> in summer is the 100% <strong>of</strong> the plant capacity, <strong>and</strong><br />

covers the 80% in winter. So, the most realistic performance data cases are (in this<br />

order) HTOS <strong>and</strong> HTOW. The error <strong>analysis</strong> is going to be underlined in these two<br />

cases.<br />

The main error source in HTOS case is detected in the pressure <strong>of</strong> the reject stages<br />

<strong>and</strong> the last stage <strong>of</strong> the recovery section (a maximum <strong>of</strong> 9% <strong>of</strong> relative error). The<br />

contractors for absolute pressure <strong>of</strong> the MSF chambers give an accuracy <strong>of</strong> two<br />

decimals, therefore the error associated to the numeric presentation could be<br />

important. The correlation to calculate the absolute pressure <strong>of</strong> a flash chamber also<br />

should improve the error detected in those values.<br />

The distillate produced in the first stages <strong>of</strong> the recovery section has a maximum<br />

relative error <strong>of</strong> 5%. This error is due to the correlations for calculating both brine <strong>and</strong><br />

steam properties <strong>and</strong> the global heat transfer coefficient <strong>of</strong> each condenser. The<br />

temperatures <strong>of</strong> the three main flows <strong>of</strong> each distiller do not exceed in any case a<br />

relative error <strong>of</strong> 1.5%.<br />

118 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Model validation<br />

5.2.2.1 NTOS case<br />

TABLE 5.23 Input data <strong>and</strong> performance parameters <strong>of</strong> the NTOS case (normal-temperature operation in<br />

summer).<br />

SW (T/h) SR (T/h) R (T/h) TBT (ºC) m ST (T/h) PR T sea (ºC)<br />

Design 19,900 17,700 19,650 100 239.6 8 32<br />

Simulator 19,965 17,396.8 19,584.5 100.4 247.9 7.85 32<br />

Rel. Error (%) –0.33 1.71 0.33 –0.40 –3.46 1.88 0.00<br />

TABLE 5.24 Model validation <strong>of</strong> the NTOS performance case.<br />

Stage<br />

p<br />

(bar)<br />

T F<br />

(ºC)<br />

T D<br />

(ºC)<br />

T B<br />

(ºC)<br />

D<br />

(T/h)<br />

p’<br />

(bar)<br />

T F’<br />

(ºC)<br />

T D’<br />

(ºC)<br />

T B’<br />

(ºC)<br />

D’<br />

(T/h)<br />

1 0.87 93.1 95.7 96.7 109 0.883 93.3 96.2 97.1 114.2 –1.49 –0.21 –0.52 –0.41 –4.77<br />

2 0.77 90 92.5 93.5 219 0.781 89.9 92.9 93.8 226.3 –1.43 0.11 –0.43 –0.32 –3.33<br />

3 0.68 86.7 89.3 90.2 328 0.688 86.6 89.5 90.4 338.9 –1.18 0.12 –0.22 –0.22 –3.32<br />

4 0.59 83.5 86 86.9 437 0.605 83.3 86.2 87.1 449.4 –2.54 0.24 –0.23 –0.23 –2.84<br />

5 0.52 80.1 82.7 83.6 545 0.531 80 82.8 83.7 557.9 –2.12 0.12 –0.12 –0.12 –2.37<br />

6 0.46 76.8 79.4 80.3 652 0.465 76.7 79.5 80.5 664.3 –1.09 0.13 –0.13 –0.25 –1.89<br />

7 0.4 73.5 76.1 77 757 0.407 73.4 76.3 77.2 768.7 –1.75 0.14 –0.26 –0.26 –1.55<br />

8 0.35 70.2 72.8 73.8 861 0.355 70.2 73 74 871 –1.43 0.00 –0.27 –0.27 –1.16<br />

9 0.3 67 69.5 70.5 963 0.309 67 69.8 70.8 971.4 –3.00 0.00 –0.43 –0.43 –0.87<br />

10 0.26 63.7 66.3 67.3 1064 0.269 63.8 66.6 67.6 1069.7 –3.46 –0.16 –0.45 –0.45 –0.54<br />

11 0.22 60.4 63 64 1160 0.234 60.6 63.5 64.5 1165.9 –6.36 –0.33 –0.79 –0.78 –0.51<br />

12 0.19 57.3 59.9 60.9 1255 0.203 57.5 60.4 61.4 1260.4 –6.84 –0.35 –0.83 –0.82 –0.43<br />

13 0.17 54.2 56.8 57.8 1347 0.175 54.4 57.3 58.3 1352.6 –2.94 –0.37 –0.88 –0.87 –0.42<br />

14 0.14 51.1 53.7 54.8 1437 0.152 51.4 54.2 55.3 1442.6 –8.57 –0.59 –0.93 –0.91 –0.39<br />

15 0.12 48 50.7 51.8 1527 0.131 48.4 51.2 52.3 1530.3 –9.17 –0.83 –0.99 –0.97 –0.22<br />

16 0.105 45 47.7 48.8 1614 0.113 45.4 48.3 49.4 1615.4 –7.62 –0.89 –1.26 –1.23 –0.09<br />

17 0.09 42.1 44.7 45.9 1698 0.097 42.5 45.4 46.6 1698 –7.78 –0.95 –1.57 –1.53 0.00<br />

18 0.08 39.5 42.5 43.7 1761 0.086 39.6 43 44.4 1763.5 –7.50 –0.25 –1.18 –1.60 –0.14<br />

19 0.07 37.1 40.2 41.5 1826 0.076 37.1 40.6 42.1 1829.7 –8.57 0.00 –1.00 –1.45 –0.20<br />

20 0.06 34.7 37.9 39.2 1896 0.067 34.6 38.2 39.7 1896.2 –11.67 0.29 –0.79 –1.28 –0.01<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 119<br />

εP<br />

(%)<br />

εT F<br />

(%)<br />

εT D<br />

(%)<br />

εT B<br />

(%)<br />

εD<br />

(%)


Simulator<br />

5.2.2.2 HTOS case<br />

TABLE 5.25 Input data <strong>and</strong> performance parameters <strong>of</strong> the HTOS case (high-temperature operation in<br />

summer).<br />

SW (T/h) SR (T/h) R (T/h) TBT (ºC) m ST (T/h) PR T sea (ºC)<br />

Design 19,900 17,700 19,850 112 294.1 8 32<br />

Simulator 19,975 17,509.1 19,850 112.3 301.3 7.86 32<br />

Rel. Error (%) –0.38 1.08 0.00 –0.27 –2.45 1.75 0.00<br />

TABLE 5.26 Model validation <strong>of</strong> the HTOS performance case.<br />

Stage<br />

p<br />

(bar)<br />

T F<br />

(ºC)<br />

T D<br />

(ºC)<br />

T B<br />

(ºC)<br />

D<br />

(T/h)<br />

p’<br />

(bar)<br />

T F’<br />

(ºC)<br />

120 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

T D’<br />

(ºC)<br />

T B’<br />

(ºC)<br />

D’<br />

(T/h)<br />

1 1.3 103.8 107.2 108.2 131 1.311 103.9 107.4 108.4 138.5 –0.85 –0.10 –0.19 –0.18 –5.73<br />

2 1.14 100.2 103.5 104.5 261 1.147 100 103.5 104.5 274.1 –0.61 0.20 0.00 0.00 –5.02<br />

3 1 96.4 99.7 100.7 390 0.997 96.2 99.5 100.5 409.9 0.30 0.21 0.20 0.20 –5.10<br />

4 0.87 92.7 95.9 96.9 521 0.865 92.2 95.6 96.6 542.8 0.57 0.54 0.31 0.31 –4.18<br />

5 0.75 88.8 92 93 649 0.748 88.3 91.7 92.7 672.9 0.27 0.56 0.33 0.32 –3.68<br />

6 0.65 85 88.2 89.2 776 0.646 84.5 87.9 88.8 800.3 0.62 0.59 0.34 0.45 –3.13<br />

7 0.56 81.1 84.4 85.3 902 0.556 80.6 84 85 924.9 0.71 0.62 0.47 0.35 –2.54<br />

8 0.48 77.2 80.5 81.5 1025 0.478 76.8 80.2 81.2 1047 0.42 0.52 0.37 0.37 –2.15<br />

9 0.41 73.4 76.7 77.7 1146 0.41 73 76.4 77.4 1166.3 0.00 0.54 0.39 0.39 –1.77<br />

10 0.35 69.6 72.9 73.9 1266 0.35 69.3 72.7 73.7 1283.3 0.00 0.43 0.27 0.27 –1.37<br />

11 0.3 65.8 69 70 1381 0.299 65.6 69 70 1397.6 0.33 0.30 0.00 0.00 –1.20<br />

12 0.25 62 65.4 66.4 1493 0.254 61.9 65.3 66.3 1509.7 –1.60 0.16 0.15 0.15 –1.12<br />

13 0.21 58.4 61.7 62.7 1603 0.215 58.3 61.7 62.7 1619.1 –2.38 0.17 0.00 0.00 –1.00<br />

14 0.18 54.7 58.1 59.1 1711 0.182 54.7 58.1 59.2 1725.8 –1.11 0.00 0.00 –0.17 –0.86<br />

15 0.15 51.1 54.5 55.5 1817 0.154 51.1 54.6 55.7 1829.7 –2.67 0.00 –0.18 –0.36 –0.70<br />

16 0.12 47.5 50.9 52 1920 0.13 47.6 51.1 52.2 1930.7 –8.33 –0.21 –0.39 –0.38 –0.56<br />

17 0.1 44 47.4 48.5 2021 0.109 44.2 47.6 48.9 2028.6 –9.00 –0.45 –0.42 –0.82 –0.38<br />

18 0.09 40.5 44.6 45.9 2096 0.095 40.9 44.9 46.2 2104.2 –5.56 –0.99 –0.67 –0.65 –0.39<br />

19 0.08 38.1 41.9 43.2 2173 0.083 38 42.2 43.6 2180.7 –3.75 0.26 –0.72 –0.93 –0.35<br />

20 0.07 35.1 39.2 40.5 2258 0.071 35 39.3 40.8 2257.9 –1.43 0.28 –0.26 –0.74 0.00<br />

εP<br />

(%)<br />

εT F<br />

(%)<br />

εT D<br />

(%)<br />

εT B<br />

(%)<br />

εD<br />

(%)


Model validation<br />

5.2.2.3 LTOS case<br />

TABLE 5.27 Some input data <strong>and</strong> performance parameters <strong>of</strong> the LTOS case (low-temperature operation in<br />

summer).<br />

SW (T/h) SR (T/h) R (T/h) TBT (ºC) m ST (T/h) PR T sea (ºC)<br />

Design 17,000 14,800 16,450 84 148.1 8.1 32<br />

Simulator 17,000 14,900.2 16,476.9 84.6 150.5 8.14 32<br />

Rel. Error (%) –0.00 –0.68 –0.16 –0.71 –1.62 –0.49 0.00<br />

TABLE 5.28 Model validation. LTOS performance case in MSF distillers.<br />

Stage<br />

p<br />

(bar)<br />

T F<br />

(ºC)<br />

T D<br />

(ºC)<br />

T B<br />

(ºC)<br />

D<br />

(T/h)<br />

p’<br />

(bar)<br />

T F’<br />

(ºC)<br />

T D’<br />

(ºC)<br />

T B’<br />

(ºC)<br />

D’<br />

(T/h)<br />

1 0.475 78.7 80.48 81.4 72 0.495 79.2 81.1 81.9 73.9 –4.21 –0.64 –0.77 –0.61 –2.64<br />

2 0.428 76.2 77.97 78.9 143 0.445 76.6 78.5 79.3 146.6 –3.97 –0.52 –0.68 –0.51 –2.52<br />

3 0.385 73.6 75.37 76.3 215 0.399 74 75.8 76.7 219.7 –3.64 –0.54 –0.57 –0.52 –2.19<br />

4 0.345 71 72.79 73.7 286 0.357 71.3 73.2 74.1 291.5 –3.48 –0.42 –0.56 –0.54 –1.92<br />

5 0.309 68.4 70.21 71.1 356 0.32 68.7 70.6 71.5 362.2 –3.56 –0.44 –0.56 –0.56 –1.74<br />

6 0.277 65.9 67.65 68 426 0.286 66.2 68 68.9 431.6 –3.25 –0.46 –0.52 –1.32 –1.31<br />

7 0.247 63.3 65.08 66 494 0.255 63.6 65.4 66.3 499.7 –3.24 –0.47 –0.49 –0.45 –1.15<br />

8 0.22 60.7 62.53 63.5 562 0.228 61.1 62.9 63.8 566.6 –3.64 –0.66 –0.59 –0.47 –0.82<br />

9 0.196 58.2 60.01 61 629 0.203 58.6 60.4 61.3 632.1 –3.57 –0.69 –0.65 –0.49 –0.49<br />

10 0.175 55.7 57.49 58.4 695 0.181 56.1 57.9 58.9 696.3 –3.43 –0.72 –0.71 –0.86 –0.19<br />

11 0.154 53.2 54.9 55.8 756 0.161 53.6 55.5 56.4 759.2 –4.55 –0.75 –1.09 –1.08 –0.42<br />

12 0.138 50.8 52.56 53.3 816 0.143 51.2 53.1 54.1 820.7 –3.62 –0.79 –1.03 –1.50 –0.58<br />

13 0.123 48.5 50.23 51.2 876 0.127 48.9 50.7 51.7 880.8 –3.25 –0.82 –0.94 –0.98 –0.55<br />

14 0.109 46.1 47.91 48.9 934 0.113 46.5 48.3 49.4 939.7 –3.67 –0.87 –0.81 –1.02 –0.61<br />

15 0.098 43.9 45.64 46.6 992 0.101 44.2 46 47.2 996.1 –3.06 –0.68 –0.79 –1.29 –0.41<br />

16 0.087 41.6 43.36 44.4 1048 0.09 42 43.7 45 1051.2 –3.45 –0.96 –0.78 –1.35 –0.31<br />

17 0.077 39.3 41.11 42.2 1104 0.08 39.8 41.5 42.8 1104.3 –3.90 –1.27 –0.95 –1.42 –0.03<br />

18 0.071 37.1 39.44 40.5 1145 0.072 37.6 39.7 41.1 1146.9 –1.41 –1.35 –0.66 –1.48 –0.17<br />

19 0.064 35.8 37.71 38.8 1187 0.066 35.8 37.9 39.4 1189.5 –3.13 0.00 –0.50 –1.55 –0.21<br />

20 0.058 33.9 35.94 37.1 1232 0.059 33.9 36 37.7 1231.9 –1.72 0.00 –0.17 –1.62 0.01<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 121<br />

εP<br />

(%)<br />

εT F<br />

(%)<br />

εT D<br />

(%)<br />

εT B<br />

(%)<br />

εD<br />

(%)


Simulator<br />

5.2.2.4 HTOW case<br />

TABLE 5.29 Some input data <strong>and</strong> performance parameters <strong>of</strong> the HTOW case (high-temperature operation in<br />

winter).<br />

SW (T/h) SR (T/h) R (T/h) TBT (ºC) m ST (T/h) PR T sea (ºC)<br />

Design 11,231.5 16,400 19,850 112 313.3 8 18<br />

Simulator 11,231 17,000 19,850 111.4 320.6 7.84 18<br />

Rel. Error (%) 0.00 –3.66 0.00 0.54 –2.33 2.00 0.00<br />

TABLE 5.30 Model validation <strong>of</strong> HTOW case <strong>of</strong> the MSF plant.<br />

Stage<br />

p<br />

(bar)<br />

T F<br />

(ºC)<br />

T D<br />

(ºC)<br />

T B<br />

(ºC)<br />

D<br />

(T/h)<br />

p’<br />

(bar)<br />

T F’<br />

(ºC)<br />

122 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

T D’<br />

(ºC)<br />

T B’<br />

(ºC)<br />

D’<br />

(T/h)<br />

1 1.28 103.2 106.8 107.9 142 1.258 102.4 106.2 107.1 148.5 1.72 0.78 0.56 0.74 –4.58<br />

2 1.12 99.3 102.8 103.8 284 1.088 98.2 102 103 294.1 2.86 1.11 0.78 0.77 –3.56<br />

3 0.97 95.2 98.7 99.7 424 0.935 94.1 97.8 98.7 439.6 3.61 1.16 0.91 1.00 –3.68<br />

4 0.83 91 94.5 95.5 565 0.801 89.8 93.5 94.5 581.9 3.49 1.32 1.06 1.05 –2.99<br />

5 0.71 86.9 90.3 91.3 703 0.684 85.7 89.4 90.3 721 3.66 1.38 1.00 1.10 –2.56<br />

6 0.6 82.7 86.2 87.2 840 0.583 81.5 85.2 86.1 857.1 2.83 1.45 1.16 1.26 –2.04<br />

7 0.51 78.5 82 83 975 0.495 77.4 81.1 82 990.1 2.94 1.40 1.10 1.20 –1.55<br />

8 0.43 74.3 77.8 78.8 1107 0.42 73.3 77 78 1120.3 2.33 1.35 1.03 1.02 –1.20<br />

9 0.36 70.1 73.7 74.7 1237 0.354 69.2 73 73.9 1247.5 1.67 1.28 0.95 1.07 –0.85<br />

10 0.3 66 69.6 70.6 1365 0.298 65.2 69 69.9 1371.7 0.67 1.21 0.86 0.99 –0.49<br />

11 0.25 61.8 65.4 66.4 1487 0.25 61.3 65 66 1493.1 0.00 0.81 0.61 0.60 –0.41<br />

12 0.21 57.8 61.4 62.5 1605 0.21 57.3 61.1 62.1 1611.8 0.00 0.87 0.49 0.64 –0.42<br />

13 0.17 53.9 57.5 58.6 1721 0.175 53.4 57.2 58.2 1727.5 –2.94 0.93 0.52 0.68 –0.38<br />

14 0.14 50 53.6 54.8 1833 0.146 49.6 53.4 54.4 1840 –4.29 0.80 0.37 0.73 –0.38<br />

15 0.12 46.2 49.8 51.1 1943 0.121 45.8 49.6 50.7 1949.2 –0.83 0.87 0.40 0.78 –0.32<br />

16 0.1 42.4 46.1 47.4 2049 0.1 42.1 45.9 47.1 2055 0.00 0.71 0.43 0.63 –0.29<br />

17 0.08 38.7 42.3 43.8 2150 0.083 38.5 42.2 43.5 2157 –3.75 0.52 0.24 0.68 –0.33<br />

18 0.07 35.2 39.5 41.1 2229 0.071 34.9 39.4 40.7 2236.8 –1.43 0.85 0.25 0.97 –0.35<br />

19 0.06 32.3 36.5 38.2 2310 0.06 31.7 36.4 37.9 2317.9 0.00 1.86 0.27 0.79 –0.34<br />

20 0.05 28.8 33.6 35.2 2400 0.051 28.4 33.3 34.9 2400 –2.00 1.39 0.89 0.85 0.00<br />

εP<br />

(%)<br />

εT F<br />

(%)<br />

εT D<br />

(%)<br />

εT B<br />

(%)<br />

εD<br />

(%)


CHAPTER 6<br />

<strong>Thermoeconomic</strong>s<br />

Fundamentals, applications <strong>of</strong> thermoeconomic diagnosis<br />

<strong>and</strong> optimization <strong>of</strong> complex energy systems<br />

As the human population grows, our finite world is becoming smaller <strong>and</strong> natural<br />

resources are more <strong>and</strong> more scarce. We must conserve them in order to survive <strong>and</strong><br />

<strong>Thermoeconomic</strong>s plays a key role in this endeavor. We should find out how energy<br />

<strong>and</strong> resources degrade, which systems work better, how to improve designs to reduce<br />

consumption <strong>and</strong> prevent residues from damaging the environment. <strong>Thermoeconomic</strong>s<br />

<strong>and</strong> its application to engineering energy systems can help to answer these<br />

questions.<br />

The production process <strong>of</strong> a complex energy system (e.g., a dual-purpose power <strong>and</strong><br />

desalination plant) can be analyzed in terms <strong>of</strong> its economic pr<strong>of</strong>itability <strong>and</strong><br />

efficiency with respect to resource consumption.<br />

An economic <strong>analysis</strong> can calculate the cost <strong>of</strong> fuel, investment, operation <strong>and</strong><br />

maintenance for the whole plant but provides no means to evaluate the single<br />

processes taking place in the subsystems nor how to distribute the costs among them.<br />

On the other h<strong>and</strong>, a thermodynamic <strong>analysis</strong> calculates the efficiencies <strong>of</strong> the<br />

subsystems <strong>and</strong> locates <strong>and</strong> quantifies the irreversibilities but cannot evaluate their<br />

significance in terms <strong>of</strong> the overall production process.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> combines economic <strong>and</strong> thermodynamic <strong>analysis</strong> by<br />

applying the concept <strong>of</strong> cost (originally an economic property) to exergy (an energetic<br />

property). Most analysts agree that exergy is the most adequate thermodynamic<br />

property to associate with cost since it contains information from the second law <strong>of</strong><br />

thermodynamics <strong>and</strong> accounts for energy quality (Tsatsaronis, 1987, 1998; Gaggioli<br />

<strong>and</strong> El-Sayed, 1987; Moran, 1990). Exergetic efficiency compares a real process to a<br />

reversible one, (i.e. an ideal process <strong>of</strong> the same type). An exergy <strong>analysis</strong> locates <strong>and</strong><br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong>s<br />

quantifies irreversibilities in a process. Exergy based thermoeconomic methods are<br />

also referred to as “exergoeconomics” (Tsatsaronis <strong>and</strong> Winhold, 1985).<br />

In his seminal book, The Entropy Law <strong>and</strong> the Economic Process,<br />

Nicholas<br />

Georgescu-Roegen (1971) pointed out that “…the science <strong>of</strong> thermodynamics began<br />

as physics <strong>of</strong> economic value <strong>and</strong>, basically, can still be regarded as such. The<br />

Entropy Law itself emerges as the most economic in nature <strong>of</strong> all natural laws… the<br />

economic process <strong>and</strong> the Entropy Law is only an aspect <strong>of</strong> a more general fact,<br />

namely, that this law is the basis <strong>of</strong> the economy <strong>of</strong> life at all levels…”.<br />

Hence, the physical magnitude connecting physics (thermodynamics) <strong>and</strong> economics<br />

is entropy generation or, more specifically, irreversibility. This represents the “useful”<br />

or available energy lost or destroyed (exergy destruction) in all physical processes.<br />

All real processes in a plant are non-reversible <strong>and</strong>, as a consequence, some exergy is<br />

destroyed <strong>and</strong> some natural resources are consumed <strong>and</strong> lost forever, which creates<br />

cost. All natural resources have an economic cost: the more irreversible a process, the<br />

more natural resources are consumed (higher energetic cost) <strong>and</strong> the higher the<br />

required investment (higher thermoeconomic cost). If we can measure this<br />

thermodynamic cost by identifying, locating <strong>and</strong> quantifying the causes <strong>of</strong><br />

inefficiencies in real processes, we can provide an objective economic basis using the<br />

cost concept.<br />

Thus, thermoeconomics assesses the cost <strong>of</strong> consumed resources, money <strong>and</strong> system<br />

irreversibilities in terms <strong>of</strong> the overall production process. Consumed resource cost<br />

involves resources destroyed by inefficiencies <strong>and</strong> helps to point out how resources<br />

may be used more effectively to save energy. Money costs express the economic<br />

effect <strong>of</strong> inefficiencies <strong>and</strong> are used to improve the cost effectiveness <strong>of</strong> production<br />

processes.<br />

Assessing the cost <strong>of</strong> the various streams <strong>and</strong> processes in a plant helps to underst<strong>and</strong><br />

the process <strong>of</strong> cost formation, from the input resource(s) to the final product(s). This<br />

process can solve problems in complex energy systems that cannot normally be<br />

solved using conventional energy <strong>analysis</strong> based on the First Law <strong>of</strong> Thermodynamics<br />

(mass <strong>and</strong> energy balances only), for instance:<br />

1. Rational price assessment <strong>of</strong> plant products based on physical criteria.<br />

2. Optimization <strong>of</strong> specific component variables to minimize final product costs <strong>and</strong><br />

save resource energy, i.e., global <strong>and</strong> local optimization.<br />

3. Detection <strong>of</strong> inefficiencies <strong>and</strong> calculation <strong>of</strong> their economic effects in operating<br />

plants, i.e., plant operation thermoeconomic diagnosis.<br />

4. Evaluation <strong>of</strong> various design alternatives or operation decisions <strong>and</strong> pr<strong>of</strong>itability<br />

maximization.<br />

5. Energy audits.<br />

124 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Specific examples <strong>of</strong> these applications will be given here <strong>and</strong> applied to a real dualpurpose<br />

power <strong>and</strong> desalination plant. Many reports also provide specific information<br />

about thermoeconomic applications (Lozano <strong>and</strong> Valero, 1993; Tsatsaronis, 1994;<br />

Lozano, Valero <strong>and</strong> Serra, 1996; Valero et al., 1994; Bejan, Tsatsaronis <strong>and</strong> Moran<br />

1997, Valero <strong>and</strong> Lozano, 1997; Valero, Correas <strong>and</strong> Serra, 1999; Lozano et al., 1994;<br />

Frangopoulos, 1987; Von Spakovsky <strong>and</strong> Evans, 1993; El-Sayed <strong>and</strong> Tribus, 1983;<br />

El-Sayed, 1988; Pisa, 1997).<br />

<strong>Thermoeconomic</strong> methods can generally be subdivided into two categories<br />

(Tsatsaronis, 1987), those based on cost accounting (e.g. Exergetic Cost Theory,<br />

Lozano et al., 1993; Average-Cost-Approach, Bejan et al., 1997; Last-In-First-Out<br />

Approach; Lazzareto <strong>and</strong> Tsatsaronis, 1997) <strong>and</strong> those based on optimization<br />

techniques (e.g. <strong>Thermoeconomic</strong> Functional Analysis, Frangopoulos, 1987;<br />

Engineering Functional Analysis, von Spakovsky <strong>and</strong> Evans, 1993; Intelligent<br />

Functional Approach, Frangopoulos, 1990). Cost accounting methods help to<br />

determine actual product cost <strong>and</strong> provide a rational basis for pricing, while<br />

optimization methods are used to find the optimum design or operating conditions.<br />

Unfortunately, there are almost as many nomenclatures as theories. This causes<br />

confusion, complicates method comparison <strong>and</strong> impedes the development <strong>of</strong><br />

<strong>Thermoeconomic</strong>s in general (Tsatsaronis, 1994). The Structural Theory <strong>of</strong><br />

<strong>Thermoeconomic</strong>s (Valero, Serra <strong>and</strong> Torres, 1992; Valero, Serra <strong>and</strong> Lozano, 1993)<br />

provides a general mathematical formulation using a linear model which<br />

encompasses all thermoeconomic methodologies. The most systematic <strong>and</strong><br />

widespread methodologies (see above) use exergy to linearly apportion costs when<br />

two or more coproducts appear, <strong>and</strong> their results can be reproduced using the<br />

Structural Theory (Erlach, 1998; Erlach, Serra <strong>and</strong> Valero, 1999). For this reason, all<br />

concepts <strong>and</strong> procedures explained here are based on the general <strong>and</strong> common<br />

mathematical formalism <strong>of</strong> the Structural Theory.<br />

This chapter on the fundamentals <strong>of</strong> thermoeconomics is divided into three parts.<br />

First the basic concepts needed to perform <strong>and</strong> underst<strong>and</strong> the thermoeconomic<br />

<strong>analysis</strong> <strong>of</strong> complex energy systems are presented. Special attention has been paid to<br />

explaining the thermoeconomic cost concept. Once the average <strong>and</strong> marginal costs<br />

are defined, in the second part their meaning, relationship <strong>and</strong> calculation procedures<br />

are fully explained with examples. Finally, the third part describes some applications<br />

<strong>of</strong> thermoeconomic <strong>analysis</strong> as applied to operation diagnosis <strong>and</strong> optimization <strong>of</strong><br />

complex energy systems.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

125


FIGURE 6.1<br />

<strong>Thermoeconomic</strong>s<br />

6.1 Basic concepts<br />

All thermoeconomic theories use costs based on the Second Law <strong>of</strong> thermodynamics<br />

when solving engineering problems. In this section, the cost concept is explained<br />

together with all the new basic concepts, including fuel, product <strong>and</strong> thermoeconomic<br />

models needed to perform a thermoeconomic <strong>analysis</strong> <strong>of</strong> a plant.<br />

Physical structure <strong>of</strong> the co-generation plant.<br />

Compressor<br />

2<br />

0<br />

6.1.1 The concept <strong>of</strong> cost<br />

2<br />

1<br />

Air<br />

Gases<br />

Natural gas<br />

Work<br />

Water/Steam<br />

Combustor<br />

1<br />

HRSG<br />

The cost <strong>of</strong> a flow in a plant represents the external resources that have to be supplied<br />

to the overall system to produce this flow. <strong>Thermoeconomic</strong> <strong>analysis</strong> distinguishes<br />

between exergetic costs <strong>and</strong> monetary costs.<br />

The exergetic cost <strong>of</strong> a mass <strong>and</strong>/or energy flow is the units <strong>of</strong> exergy used to produce<br />

it, e.g. the exergetic cost <strong>of</strong> the net power is the exergy provided by the natural gas to<br />

generate the electrical power delivered to the net by the cogeneration plant (see figure<br />

6.1). These costs are a measure <strong>of</strong> the thermodynamic efficiency <strong>of</strong> the production<br />

process generating these flows. The unit exergetic cost <strong>of</strong> a mass <strong>and</strong>/or energy flow<br />

represents the amount <strong>of</strong> resources required to obtain one unit. Thus, if the unit<br />

exergetic cost <strong>of</strong> the electricity is three, three units <strong>of</strong> plant exergy resources (natural<br />

gas in the case <strong>of</strong> the cogeneration plant) are consumed to obtain one exergy unit <strong>of</strong><br />

electrical power.<br />

The monetary cost takes into account the economic cost <strong>of</strong> the consumed fuel (i.e., its<br />

market price) as well as the cost <strong>of</strong> the installation <strong>and</strong> the operation <strong>of</strong> the plant <strong>and</strong><br />

defines the amount <strong>of</strong> money consumed to generate a mass <strong>and</strong>/or energy flow. These<br />

costs are a measure <strong>of</strong> the economic efficiency <strong>of</strong> a process. Similarly, the unit<br />

126 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

3<br />

5 6<br />

Turbine<br />

3<br />

4<br />

8<br />

4<br />

7


Basic concepts<br />

monetary cost (also called unit exergoeconomic cost or unit thermoeconomic cost)<br />

<strong>of</strong> a<br />

mass <strong>and</strong>/or energy flow is the amount <strong>of</strong> monetary units required to obtain one unit.<br />

We can further distinguish between average costs,<br />

which are ratios <strong>and</strong> express the<br />

average amount <strong>of</strong> resources per unit <strong>of</strong> product, <strong>and</strong> marginal costs,<br />

which are a<br />

derivation <strong>and</strong> indicate the additional resources required to generate one more unit <strong>of</strong><br />

the product under specified conditions. Mathematically they are defined as:<br />

unit average cost = -----<br />

(6.1)<br />

⎛ )<br />

* ∂B<br />

⎞<br />

o<br />

unit marginal cost k = ⎜<br />

(6.2)<br />

⎜<br />

⎟<br />

⎝ ∂B<br />

⎟<br />

i ⎠<br />

The average costs are only known after production, when we know how many<br />

resources were used <strong>and</strong> the production obtained. The average cost is not predictive.<br />

Knowing the average unit cost <strong>of</strong> a product does not provide the cost <strong>of</strong> a production<br />

process P + ∆P.<br />

<strong>Thermoeconomic</strong> cost accounting theories calculate average costs<br />

<strong>and</strong> use them as a basis for a rational price assessment, under physical criteria, <strong>of</strong> the<br />

internal flows <strong>and</strong> the products <strong>of</strong> the plant.<br />

Marginal costs can be used to calculate additional fuel consumption when the<br />

operating conditions are modified. <strong>Thermoeconomic</strong> optimization methods<br />

(Frangopoulos, 1997, 1990; Von Spakovsky <strong>and</strong> Evans, 1993) are based on marginal<br />

costs when solving optimization problems.<br />

The relationship between average <strong>and</strong> marginal costs will be analyzed in more detail<br />

in section 6.3.1.<br />

6.1.2 Fuel, product <strong>and</strong> unit exergetic consumption<br />

k *<br />

B0 Bi conditions<br />

A productive purpose, a certain good or service to be produced, can be defined for<br />

every plant. In order to generate this product, some resources have to be supplied to<br />

the plant <strong>and</strong> are consumed in the process. For example, in the co-generation plant,<br />

natural gas is supplied to the plant to generate electric power <strong>and</strong> process steam.<br />

A productive purpose expressing component function in an overall production<br />

process can be defined for each component. The productive purpose <strong>of</strong> a component<br />

measured in terms <strong>of</strong> exergy is called product.<br />

To create this product, another exergy<br />

flow(s) is consumed. The flow <strong>of</strong> exergy which is consumed in the component during<br />

the generation <strong>of</strong> its product is called fuel (s) .<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

127


<strong>Thermoeconomic</strong>s<br />

Real process exergy is destroyed in any process. That is, part <strong>of</strong> the fuel exergy is<br />

destroyed during product generation. Using the definitions <strong>of</strong> fuel <strong>and</strong> product, the<br />

exergy balance for a component can be formulated as:<br />

F = P + I (6.3)<br />

Therefore, the fuel required to generate a certain amount <strong>of</strong> a product depends on the<br />

amount <strong>of</strong> irreversibility (exergy destroyed).<br />

The fuel exergy required to generate one exergy unit <strong>of</strong> product is defined as unit<br />

exergetic consumption k:<br />

k<br />

=<br />

F<br />

--<br />

P<br />

128 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(6.4)<br />

It is a measure <strong>of</strong> the thermodynamic efficiency <strong>of</strong> the process <strong>and</strong> equals one for<br />

reversible processes <strong>and</strong> is greater than one for all real processes. The more<br />

irreversible a process, the higher the value <strong>of</strong> the unit exergetic consumption.<br />

Combining equation (6.4) with the exergy balance on a fuel/product basis (Equation<br />

6.3), the unit exergetic consumption k can also be formulated as:<br />

I<br />

k = 1+<br />

--<br />

P<br />

(6.5)<br />

The reciprocal <strong>of</strong> the unit exergy consumption is defined as the exergetic efficiency η.<br />

It is equal to one for reversible processes <strong>and</strong> is less than one for all real processes.<br />

P I<br />

η = -- =<br />

1–<br />

--<br />

F F<br />

(6.6)<br />

Fuel <strong>and</strong> product definitions for some typical components in a dual-purpose power<br />

<strong>and</strong> desalination plant are shown in table 6.1. The fuel-product definition for the<br />

components <strong>of</strong> the cogeneration plant (figure 6.1) are shown in table 6.2.


TABLE 6.1<br />

Basic concepts<br />

Fuel <strong>and</strong> product definitions for typical dual-purpose power <strong>and</strong> desalination plant units.<br />

Component Fuel Product<br />

Boiler<br />

Pump<br />

Turbine<br />

without<br />

extraction<br />

Turbine<br />

with<br />

extraction<br />

Generator<br />

Heat<br />

exchanger/<br />

brine heater<br />

MSF stage<br />

B 1<br />

fuel<br />

B 1<br />

W mech<br />

B 1<br />

B 1<br />

B 1<br />

cold<br />

stream<br />

B 4<br />

B 1<br />

W<br />

B 2<br />

B 2<br />

B 2 water<br />

B 3<br />

steam<br />

B 3<br />

W<br />

W<br />

B 2<br />

W el<br />

B4 B2 B 3<br />

hot stream<br />

D<br />

B 3<br />

B 2<br />

Natural gas<br />

B<br />

1<br />

Work to drive pump/compressor<br />

W<br />

Exergy removed from working<br />

fluid during the expansion<br />

B1<br />

– B2<br />

Exergy removed from working<br />

fluid during the expansion<br />

B1<br />

– B2<br />

– B3<br />

Mechanical work<br />

W<br />

mech<br />

Exergy removed from the hot<br />

flow<br />

B3<br />

– B4<br />

Exergy removed from the<br />

flashing brine (B1<br />

– B2)<br />

minus<br />

exergy provided to the cooling<br />

brine (B4<br />

– B3)<br />

Exergy difference between the<br />

generated steam flow <strong>and</strong> the<br />

entering water flow<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

B<br />

3<br />

– B<br />

2<br />

Exergy supplied to the working fluid<br />

B<br />

2<br />

– B<br />

1<br />

Generated work<br />

W<br />

Generated work<br />

W<br />

Electric Work<br />

W<br />

el<br />

Exergy supplied to the cold flow<br />

B2<br />

– B1<br />

Distilled water in the stage<br />

D<br />

129


FIGURE 6.2<br />

<strong>Thermoeconomic</strong>s<br />

6.1.3 Physical <strong>and</strong> thermoeconomic plant models<br />

A plant is analyzed using a physical model with a group <strong>of</strong> equations to describe the<br />

physical behavior <strong>of</strong> the components. It calculates parameters such as temperatures,<br />

pressures, efficiencies, power generated etc. to describe the physical state <strong>of</strong> the plant.<br />

Depending on the <strong>analysis</strong>, a decision has to be taken on the detail required i.e.,<br />

which flows <strong>and</strong> components are to be considered. The components for the <strong>analysis</strong><br />

do not necessarily correspond to physical units. Various parts <strong>of</strong> the installation can<br />

be <strong>combined</strong> into one component <strong>and</strong> physical units can be further disaggregated. It<br />

is important to chose an appropriate aggregation level that properly defines the<br />

behavior <strong>of</strong> each component <strong>and</strong> its purpose in the overall production process. The<br />

physical structure (see figure 6.1) depicts the components, mass <strong>and</strong> connecting<br />

energy flows considered in the physical model.<br />

The minimum physical data required in a thermoeconomic <strong>analysis</strong> are temperatures,<br />

pressures, mass flow rates <strong>and</strong> compositions <strong>of</strong> all mass flows together with the heat<br />

<strong>and</strong> power rates <strong>of</strong> the energy flows considered. Usually all this information is fully<br />

or partially obtained from the physical model <strong>of</strong> the plant. But it is not strictly<br />

indispensable if all the required data are measured plant data, collected directly from<br />

the plant data acquisition system.<br />

Productive structure <strong>of</strong> the cogeneration plant.<br />

Compressor<br />

P 2 = B 2 – B 0<br />

P 1 = B 3 – B 2<br />

F 1 = B 1<br />

Combustor<br />

Pj1 = B3 j1<br />

F 3 = B 3 – B 4<br />

F 2 = B 5 = W Cp<br />

F 4 = B 4<br />

Turbine<br />

HRSG<br />

P 4 = B 7 = B heat<br />

Nevertheless, when pricing all mass <strong>and</strong> energy flows in the thermoeconomic<br />

<strong>analysis</strong>, it is absolutely necessary to define a thermoeconomic model <strong>of</strong> the plant<br />

which considers the productive purpose <strong>of</strong> the components, i.e. the definitions <strong>of</strong><br />

fuels <strong>and</strong> products <strong>and</strong> the distribution <strong>of</strong> the resources throughout the plant. The<br />

productive model can be graphically depicted by the productive structure diagram<br />

(figure 6.2).<br />

130 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

b1<br />

P 3<br />

b2<br />

W net = B 6


TABLE 6.2<br />

Basic concepts<br />

In this scheme, the flows (lines connecting the equipment) are the fuel <strong>and</strong> the<br />

product <strong>of</strong> each subsystem. Each “real“ piece <strong>of</strong> equipment in the plant has an outlet<br />

flow (product) <strong>and</strong> an inlet flow (fuel). The capital cost <strong>of</strong> the units is also considered<br />

as an external plant resource <strong>and</strong> is represented as inlet flows coming directly from<br />

the environment (not considered in figure 6.2). Since the fuel <strong>of</strong> a process unit can be<br />

the product <strong>of</strong> another <strong>and</strong> the product <strong>of</strong> a process unit can be the fuel <strong>of</strong> several<br />

subsystems, two types <strong>of</strong> fictitious devices are introduced: junctions (rhombs) <strong>and</strong><br />

branching points or branches (circles). In a junction, the products <strong>of</strong> two or more<br />

components are joined to form the fuel <strong>of</strong> another component. In a branching point,<br />

an exergy flow (fuel or product in the productive structure –see figure 6.2-) is<br />

distributed between two or more components. Sometimes the productive structure<br />

can be simplified (with the same results) by merging the junctions <strong>and</strong> branches in a<br />

new fictitious component called junction-branching point. Figure 6.5 in section 6.3.1<br />

shows a similar productive structure as figure 6.2, where the junction j1 <strong>and</strong> the<br />

branching point b1 have been merged in a junction-branching point. For the sake <strong>of</strong><br />

simplicity, the explanation <strong>of</strong> the fundamentals <strong>of</strong> thermoeconomics will be made<br />

using the productive structure depicted in figure 6.2.<br />

Fuels <strong>and</strong> Products <strong>of</strong> the components <strong>of</strong> the co-generation plant.<br />

No Subsystem Fuel Product<br />

Technical<br />

production<br />

coefficients<br />

1 Combustor F<br />

1 = B 1 P 1 = B 3 – B 2 k cb = F 1 /P 1<br />

2 Compressor F 2 = B 5 = W cp P 2 = B 2 – B 0 k cp = F 2 /P 2<br />

3 Turbine F 3 = B 3 – B 4 P 3 = B 5 + B 6 = W cp + W net k gt = F 3 /P 3<br />

4 HRSG F 4 = B 4 P 4 = B 7 = B heat k HRSG = F 4/P 4<br />

5 Junction<br />

P 1 = B 3 – B 2<br />

P 2 = B 2 – B 0<br />

6 Branching 1 P j1 = B 3<br />

7 Branching 2 P 3 = B 5 + B 6 = W cp + W net<br />

P j1 = B 3<br />

F 3 = B 3 – B 4<br />

F 4 = B 4<br />

F 2 = B 5 = W cp<br />

B 6 = W net<br />

r 1 = P 1/P j1<br />

r 2 = P 2 /P j1<br />

The productive structure is a graphical representation <strong>of</strong> resource distribution<br />

throughout the plant. Thus, its flows are fictitious <strong>and</strong> are not necessarily physical<br />

flows. While each plant has only one physical structure to describe the physical<br />

relations between the components, various productive structures can be defined<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

131


<strong>Thermoeconomic</strong>s<br />

depending on the fuel <strong>and</strong> product definitions as well as decisions on how the plant<br />

resources are distributed among the components.<br />

Thus, the thermoeconomic model (mathematical representation <strong>of</strong> the productive<br />

structure) is a set <strong>of</strong> mathematical functions called characteristic equations, which<br />

express each inlet flow as a mathematical function <strong>of</strong> the outlet flows for all the<br />

productive structure components <strong>and</strong> a set <strong>of</strong> internal parameters x l:<br />

B i = g i (x l, B j) i = 1,…, m–s (6.7)<br />

where the index i refers to the input flows <strong>of</strong> the component l, the index j refers to the<br />

output flows <strong>of</strong> the component l, <strong>and</strong> m is the number <strong>of</strong> flows considered in the<br />

productive structure. Every flow is an input flow <strong>of</strong> a component <strong>and</strong> an output flow<br />

<strong>of</strong> another component or the environment. For the flows interacting with the<br />

environment, we define:<br />

B m–s+1 = ω i i = 1,…, s (6.8)<br />

where s is the number <strong>of</strong> system outputs, <strong>and</strong> ω i is the total system product, i.e., an<br />

external variable which determines the total product. The characteristic equations for<br />

the system in figure 6.2, are shown in table 6.3:<br />

TABLE 6.3 Characteristic equations a <strong>of</strong> the cogeneration plant.<br />

No Component Entry Outlet Equation<br />

1 Combustor F 1 P 1 F 1 = g F1 (x 1 , P 1 ) = k cb P 1<br />

2 Compressor F 2 = W cp P 2 F 2 = g F2 (x 2 , P 2 ) = k cp P 2<br />

3 Turbine F 3 P 3 = W gt F 3 = g F3 (x 3 , P 3 ) = k gt P 3<br />

4 HRSG F 4 P 4 = B heat = ω 4<br />

5 Junction 1 P 1 , P 2 P j1<br />

F 4 = g F4 (x 4 , P 4 ) = k HRSG P 4 = k HRSG ω 4<br />

= k HRSG B heat<br />

P1 = gP1 (x5 , Pj1 ) = r1 Pj1 = r1 (F3 + F4 )<br />

P2 = gP2 (x5 , Pj1 ) = r2 Pj1 = r2 (F3 + F4 )<br />

6 Branching 1 P j1 F 3 , F 4 P j1 = g Pj1 (x 6 , F 3 , F 4 ) = (F 3 + F 4 )<br />

7 Branching 2 P 3 F 2 , W net<br />

P 3 = g P3 (x 7 , F 2 , ω 3 ) = F 2 + ω 3<br />

= W cp + W net<br />

a. Variables in these characteristic equations are from the Exergetic Cost Theory, corresponding to the PF representation<br />

(Torres, 1991, Valero <strong>and</strong> Torres, 1990). Note that the Exergetic Cost Theory is a particular case<br />

(Serra, 1994) <strong>of</strong> the Structural Theory which is the thermoeconomic mathematical formalism presented in<br />

this Ph. D. Thesis.<br />

132 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Basic concepts<br />

The inlet <strong>and</strong> outlet flows <strong>of</strong> the productive structure units are extensive magnitudes,<br />

which are the product <strong>of</strong> a quantity (usually mass flow rate) <strong>and</strong> a quality (specific<br />

magnitude). The magnitudes applied by most thermoeconomic methodologies are<br />

exergy (Tsatsaronis, 1987), negentropy (Frangopoulos, 1983) <strong>and</strong> money. Other<br />

magnitudes, like enthalpy or entropy, can also be used.<br />

The internal variables appearing in the thermoeconomic model depend on the<br />

behavior <strong>of</strong> the subsystem <strong>and</strong> they are presumably independent <strong>of</strong> mass flow rates.<br />

This implies that relations like efficiencies or pressure <strong>and</strong> temperature ratios —<br />

which are mainly independent <strong>of</strong> the quantity <strong>of</strong> the exiting flows— can be used as<br />

internal parameters.<br />

Note, that the main objective <strong>of</strong> the productive structure, <strong>and</strong> hence <strong>of</strong> the<br />

thermoeconomic model, consists on sorting the thermodynamic magnitudes related<br />

to the physical mass <strong>and</strong> energy flow-streams connecting the plant subsystems, in a<br />

different way that the equations modeling the physical plant behavior do, in order to<br />

explicitly determine for each subsystem its energy conversion efficiency.<br />

It is important take in mind that, as it was already explained, thermoeconomics<br />

connects thermodynamics, which is a phenomenological (black box <strong>analysis</strong>)<br />

science, with economics. That is, by sorting the thermodynamic properties <strong>of</strong> the<br />

physical mass <strong>and</strong> energy flow-streams <strong>of</strong> a plant, which in turn provide the energy<br />

conversion efficiency <strong>of</strong> each subsystem, thermoeconomics analyzes the degradation<br />

process <strong>of</strong> energy quality through an installation, i.e., thermoeconomics evaluates the<br />

process <strong>of</strong> cost formation.<br />

Depending on the <strong>analysis</strong> scope each subsystem can be identified with a separate<br />

piece <strong>of</strong> equipment, a part <strong>of</strong> a device, several process units or even the whole plant.<br />

Sometimes the objective consists on analyzing a plant in a deep detail. In this case it<br />

is advisable, if possible, to identify each subsystem with a separate physical process<br />

(heat transfer, pressure increase or decrease <strong>and</strong> chemical mixture or reaction) in<br />

order to locate <strong>and</strong> quantify, separately if possible, each thermal, mechanical <strong>and</strong><br />

chemical irreversibility occurring in the plant. If the objective consists on analyzing a<br />

macro-system composed <strong>of</strong> several plants, probably in this case the more convenient<br />

approach is consider each separate plant as a subsystem.<br />

Thus, thermoeconomics always performs a systemic <strong>analysis</strong>, no matter how<br />

complex the system is, basically oriented to locate <strong>and</strong> quantify the energy<br />

conversion efficiency. It is out <strong>of</strong> the scope <strong>of</strong> thermoeconomics to model the<br />

behavior <strong>of</strong> the components, which is made by the mathematical equations <strong>of</strong> the<br />

physical model.<br />

Even though (it is out <strong>of</strong> the scope <strong>of</strong> thermoeconomics simulate the behavior <strong>of</strong> the<br />

subsystems), it is very important build the thermoeconomic model with physical<br />

meaning. This is the reason, as already explained, <strong>of</strong> defining different<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 133


<strong>Thermoeconomic</strong>s<br />

thermoeconomic models for the same plant. Depending on the aggregation level <strong>and</strong><br />

on the nature <strong>of</strong> the thermoeconomic equations the model will content physical<br />

information about the actual system behavior with different accuracy degrees. The<br />

obtained results from a very rough thermoeconomic model, without any physical<br />

sensitivity related with the actual behavior <strong>of</strong> the plant, probably will be useless.<br />

The more extended thermoeconomic methodologies use linear equations in their<br />

thermoeconomic models, because they present practical (the model is simpler <strong>and</strong> for<br />

this reason much more powerful when applied to very complex energy systems) <strong>and</strong><br />

conceptual advantages, as it will be explained before. Moreover, in many real plants it<br />

is possible to find an aggregation level where the system <strong>and</strong> subsystems linearly<br />

behave with accuracy enough, under an engineering point <strong>of</strong> view (Valero, Torres <strong>and</strong><br />

Lerch, 1999; Martínez, Serra <strong>and</strong> Valero, 2000). This is also the case <strong>of</strong> the dual<br />

power <strong>and</strong> desalination plant analyzed in this work, as it is proved in next chapter.<br />

Thus, if the characteristic equations are first grade homogeneous functions with<br />

respect to the subset B, <strong>of</strong> independent variables (as linear equations do), that is:<br />

λ B i = g i (λ B 1,… λ B j, x l) λ∈ℜ (6.9)<br />

Euler’s Theorem states that the homogeneous function <strong>of</strong> first order verify:<br />

gi<br />

gi<br />

gi<br />

Bi<br />

= Bl<br />

Bl<br />

Bl<br />

Bl<br />

Bl<br />

Bls<br />

∂<br />

⎛ ⎞<br />

⎜ ⎟ +<br />

⎜ ∂ ⎟<br />

⎝ ⎠<br />

∂<br />

⎛ ⎞<br />

⎜ ⎟ + +<br />

⎜ ∂ ⎟<br />

⎝ ⎠<br />

∂<br />

⎛ ⎞<br />

... ⎜ ⎟<br />

1<br />

2 ⎜ ∂ ⎟<br />

1<br />

2 ⎝ ⎠<br />

or using the marginal consumption notation,<br />

κ ij<br />

gi<br />

=<br />

B<br />

∂<br />

∂<br />

B = ∑κ B<br />

i ij j<br />

j∈Sl j<br />

l 1,…,l s in S l<br />

134 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(6.10)<br />

(6.11a)<br />

i = 1,...,m l = 1,...,n. (6.11b)<br />

This property means that the input <strong>of</strong> a component varies at the same rate as its<br />

outputs. Note that this property does not imply that the function must be linear. For<br />

instance, a Cobb-Douglas function z = a x α y (1–α) , is also a homogeneous first order<br />

function.<br />

κ ij are the technical production coefficients <strong>and</strong> represent the portion <strong>of</strong> the i-th<br />

component production:<br />

κ ij<br />

gi<br />

=<br />

B<br />

∂<br />

∂<br />

j<br />

s<br />

(6.12a)


Basic concepts<br />

The sum <strong>of</strong> κ ij coefficients <strong>of</strong> a unit is the unit exergy consumption <strong>of</strong> that unit:<br />

k<br />

Fi<br />

n<br />

i=<br />

0 Fj<br />

= ∑ κ = =<br />

P P<br />

j ij<br />

i=<br />

0<br />

n<br />

∑<br />

j<br />

j<br />

(6.12b)<br />

In thermoeconomics there are three types <strong>of</strong> characteristic equations, which are<br />

linear:<br />

1. Those connecting each fuel <strong>of</strong> a component to its corresponding product:<br />

F i = κ ij P j as for instance F 1 = g F1 (x 1, P 1) = k cb P 1<br />

(6.13a)<br />

There is one such equation for each component’s fuel. These types <strong>of</strong> equations<br />

are generated in the pieces <strong>of</strong> equipment <strong>and</strong> they inform about:<br />

– the productive function <strong>of</strong> each component, i.e., its production (product)<br />

– what the component needs (fuel) to develop its productive purpose, <strong>and</strong><br />

– the thermodynamic efficiency <strong>of</strong> the process in the component<br />

2. Structural equations model how the resources consumed by the plant are distributed<br />

through the plant components. They show how the process units are connected<br />

from a productive point <strong>of</strong> view. Structural equations are characteristic<br />

equations to describe the productive model <strong>of</strong> junctions <strong>and</strong> branches, e.g.:<br />

P 1 = g P1 (x 5, P j1) = r 1 P j1 = r 1 (F 3+F 4) (6.13b)<br />

3. When the capital cost <strong>of</strong> the equipment is also considered in the <strong>analysis</strong>, a third<br />

type <strong>of</strong> characteristic equation is required; costing equations. These equations are<br />

very <strong>of</strong>ten not linear, but in the case <strong>of</strong> these equations this is a minor problem,<br />

because they can be linearized for different operation intervals. They relate the<br />

investment cost <strong>of</strong> the component with thermodynamic variables <strong>and</strong> its product.<br />

They express the amount <strong>of</strong> resources needed to build, install, maintain (etc.) a<br />

component. For example, a costing equation proposed by El-Sayed (1996), see<br />

section 7.3.3.1 for details:<br />

Z = 002 . ⋅10⋅Q⋅∆T ⋅∆T ⋅∆P<br />

−075 . −05 . −01<br />

.<br />

n t t<br />

(6.13c)<br />

The diagram <strong>of</strong> the productive structure is also called a Fuel/Product diagram (Torres<br />

et al., 1999) because in most cases the lines connecting the pieces <strong>of</strong> equipment<br />

represent the fuels <strong>and</strong> products <strong>of</strong> the different units. Thus, the characteristic<br />

equations (see table 6.3) using the Fuel–Product notation can also be written as:<br />

P = B + B<br />

i i0 ij<br />

j=<br />

1<br />

n<br />

∑<br />

i = 0,1,…, n (6.14)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 135


<strong>Thermoeconomic</strong>s<br />

This equation shows how the production <strong>of</strong> a process unit is used as fuel by another<br />

unit or as a part <strong>of</strong> the total plant production. In the above expression, B ij is the<br />

production portion <strong>of</strong> the i-th component that fuels the j-th component, <strong>and</strong> B i0<br />

represents the production portion <strong>of</strong> the component i leading to the final plant product<br />

(the subscript 0 refers to the environment, which is considered another process unit<br />

interacting with the plant).<br />

Equation (6.14) can be expressed in terms <strong>of</strong> the unit exergetic consumptions as:<br />

P = B + κ P<br />

i i0 ij j<br />

j=<br />

1<br />

n<br />

∑<br />

In matrix notation it can also be expressed as:<br />

P = Ps + KP P<br />

i = 0,1,…, n (6.15)<br />

136 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(6.16)<br />

where P s is a (n×1) vector whose elements contain the contribution to the final<br />

production <strong>of</strong> the system P i0 obtained in each component, <strong>and</strong> 〈KP〉 is a (n×n) matrix,<br />

whose elements are the unit exergy consumption κ ij. This expression helps to relate<br />

the production <strong>of</strong> each component as a function <strong>of</strong> the final production <strong>and</strong> the unit<br />

consumption <strong>of</strong> each component:<br />

⎛ ⎞<br />

P = P Pswhere<br />

P ≡⎜UD− KP ⎟<br />

(6.17)<br />

⎝ ⎠<br />

In the same way, we can express the irreversibility <strong>of</strong> each component as:<br />

( D D)<br />

I = I Pswhere<br />

I ≡ K −U<br />

P<br />

(6.18)<br />

while the total resources <strong>of</strong> the system may be obtained as:<br />

t<br />

t<br />

FT =κe P Ps<br />

(6.19)<br />

where κe<br />

≡<br />

( κ01, …κ , 0n)<br />

, is a (n×1) vector whose elements contain the unit<br />

consumption <strong>of</strong> the system-input resources.<br />

6.2 Calculating thermoeconomic costs<br />

Once the thermoeconomic model has been defined <strong>and</strong> the characteristic equations<br />

corresponding to the productive structure <strong>of</strong> the system are known, the costs <strong>of</strong> all<br />

flows in the productive structure can be easily calculated.<br />

There are two different types <strong>of</strong> thermoeconomic costs: average costs <strong>and</strong> marginal<br />

costs (equations 6.1 <strong>and</strong> 6.2). It is important to note that (as discussed below) the<br />

−1


Calculating thermoeconomic costs<br />

average <strong>and</strong> marginal costs coincide when the characteristic equations <strong>of</strong> the<br />

thermoeconomic model are first grade homogeneous functions.<br />

This result is very important since both costs can be calculated using the same<br />

procedure. Marginal costs are a derivative (see equation 6.2) <strong>and</strong> can be calculated by<br />

applying the chain rule <strong>of</strong> the mathematical derivation. Similarly, average costs can<br />

also be obtained from the rules <strong>of</strong> the mathematical derivation applied to the<br />

thermoeconomic model when the characteristic equations are first grade<br />

homogeneous functions.<br />

According to the previous premises, the cost <strong>of</strong> the plant resources can be defined as:<br />

e<br />

0 = ∑ 0,<br />

i<br />

i=<br />

1<br />

*<br />

B k B<br />

i<br />

(6.20)<br />

where e, is the number <strong>of</strong> system inputs, <strong>and</strong> k * 0,i is the unit cost <strong>of</strong> the –i– external<br />

resource.<br />

Each flow, as a component input, is a function (defined by the characteristic equation) <strong>of</strong> a<br />

set <strong>of</strong> internal variables, x, external variables ω <strong>and</strong> the output flows <strong>of</strong> the component.<br />

The cost <strong>of</strong> the plant resources is then a function <strong>of</strong> each flow, the set <strong>of</strong> internal<br />

variables <strong>of</strong> each component <strong>and</strong> the final product <strong>of</strong> the plant B 0 = B 0 (B i , x, ω),<br />

according the relations (6.7) <strong>and</strong> (6.8).<br />

When calculating the variation <strong>of</strong> the resources consumed in the plant concerning a<br />

flow, the chain rule can be applied:<br />

∂B<br />

∂B<br />

0<br />

i<br />

∂B 0<br />

---------<br />

∂B i<br />

= k<br />

=<br />

*<br />

0, i<br />

m<br />

∑<br />

j = 1<br />

∂B 0<br />

i = 1,…, e (6.21a)<br />

∂B 0<br />

--------- ∂g j<br />

--------<br />

∂B j<br />

∂B i<br />

i = e + 1,…, m (6.21b)<br />

The expression --------- represents the marginal costs which evaluate the additional<br />

∂B i<br />

consumption <strong>of</strong> the resources, when an additional unit <strong>of</strong> the flow –i– is produced,<br />

under the conditions that the internal variables, x, do not vary throughout this process.<br />

We can denote these marginal costs as k* i , <strong>and</strong> κij = -------- the marginal consumption<br />

<strong>of</strong> flow –i– to produce the flow –j–, then we can rewrite the previous expressions, as:<br />

*<br />

i = 0,<br />

i<br />

*<br />

k k<br />

i = 1,…, e (6.22a)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 137<br />

∂g i<br />

∂B j


<strong>Thermoeconomic</strong>s<br />

i = e + 1,…, m (6.22b)<br />

Note that the unit exergetic cost <strong>of</strong> each fuel entering the plant is unity because there<br />

is no energy quality degradation nor exergy destruction at the very beginning <strong>of</strong> the<br />

productive process. Hence, the amount <strong>of</strong> exergy consumed to obtain each plant’s<br />

fuel is its own exergy content <strong>and</strong> therefore its unit exergetic cost equals one.<br />

It can easily be proved that the cost <strong>of</strong> each flow Pij <strong>of</strong> the productive structure using<br />

the Fuel/Product notation is:<br />

138 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(6.23)<br />

And the exergetic cost <strong>of</strong> the product <strong>of</strong> each component is the same as the cost <strong>of</strong> the<br />

resources needed to obtain it, hence:<br />

i = 1,…, n (6.24)<br />

This cost equation can also be expressed in terms <strong>of</strong> the unit exergetic consumptions:<br />

i = 1,…, n (6.25)<br />

which can be used to obtain the unit exergetic cost <strong>of</strong> the flows appearing in the<br />

productive structure diagram as a function <strong>of</strong> the unit exergetic consumption <strong>of</strong> each<br />

process unit.<br />

Then, if the characteristic equations <strong>and</strong> the marginal consumptions for each<br />

component are known, the marginal cost k * for each flow can be obtained by solving<br />

the system <strong>of</strong> linear equations (6.25).<br />

Example 1<br />

m<br />

∑<br />

k* *<br />

i = κ ji k j<br />

j = 1<br />

j≠i * *<br />

P = k B<br />

ij P, i ij<br />

n<br />

i i ∑ P j ji<br />

0<br />

* *<br />

*<br />

P = F = k , B<br />

j=<br />

n<br />

Pi i ji<br />

*<br />

Pj ,<br />

∑<br />

*<br />

k , = κ0+ κ k<br />

j=<br />

1<br />

For the example <strong>of</strong> a co-generation plant (figure 6.2), equations 6.21a, 6.21b can be<br />

written as:<br />

* ∂B<br />

k F = 1 ∂F<br />

k<br />

k<br />

1<br />

1<br />

B<br />

=<br />

F<br />

∂<br />

∂<br />

B<br />

=<br />

P<br />

∂<br />

∂<br />

∂P<br />

∂F<br />

= k<br />

* 1 1 3 *<br />

F2 P3<br />

2 3 2<br />

B B P<br />

1 1 j1<br />

= k<br />

F P F<br />

∂<br />

=<br />

∂<br />

∂ ∂<br />

=<br />

∂ ∂<br />

* *<br />

F3<br />

Pj1 3 j1<br />

3<br />

*


Calculating thermoeconomic costs<br />

k<br />

k<br />

k<br />

k<br />

k<br />

k<br />

k<br />

B<br />

=<br />

F<br />

∂<br />

∂<br />

B<br />

=<br />

P<br />

∂<br />

∂<br />

∂Pj<br />

∂F<br />

= k<br />

* 1 1 1 *<br />

F4<br />

Pj1 4 j1<br />

4<br />

B<br />

=<br />

P<br />

∂<br />

∂<br />

B<br />

=<br />

F<br />

∂<br />

∂<br />

∂F1<br />

k k<br />

∂ P<br />

=<br />

* 1 1<br />

*<br />

P1 F1 cb<br />

1 1 1<br />

B<br />

=<br />

P<br />

∂<br />

∂<br />

B<br />

=<br />

F<br />

∂<br />

∂<br />

∂F<br />

∂P<br />

= k k<br />

* 1 1 2 *<br />

P2 F2 cp<br />

2 2 2<br />

B<br />

=<br />

P<br />

∂<br />

∂<br />

B<br />

=<br />

F<br />

∂<br />

∂<br />

∂F<br />

∂P<br />

= k k<br />

* 1 1 3 *<br />

P3 F3 gt<br />

3 3 3<br />

B<br />

=<br />

P<br />

∂<br />

∂<br />

B<br />

=<br />

F<br />

∂<br />

∂<br />

∂F<br />

∂P<br />

= k k<br />

* 1 1 4 *<br />

P4 F4 HRSG<br />

4 4 4<br />

B<br />

=<br />

P<br />

∂<br />

∂<br />

B<br />

=<br />

P<br />

∂<br />

∂<br />

∂P<br />

∂P<br />

B<br />

+<br />

P<br />

∂<br />

∂<br />

∂P<br />

∂P<br />

= k r + k r<br />

* 1 1 2 1 1 * *<br />

Pj1<br />

P2 2 P1<br />

1<br />

j12j11j1 B<br />

=<br />

W<br />

∂<br />

∂<br />

B<br />

=<br />

P<br />

∂<br />

∂<br />

∂P<br />

∂W<br />

= k<br />

* 1 1 3 *<br />

Wnet<br />

P3<br />

net 3 net<br />

The thermoeconomic model (characteristic equations) <strong>of</strong> an energy system contains<br />

the mathematical dependence between the resources consumed <strong>and</strong> plant flows<br />

(products <strong>and</strong> internal flows). It is therefore possible to define a set <strong>of</strong> linear equations<br />

to calculate the costs <strong>of</strong> every flow <strong>of</strong> the plant's productive structure. Note that these<br />

equations show the process <strong>of</strong> cost formation on the productive structure.<br />

The proposed procedure to calculate the marginal cost <strong>of</strong> all the flows <strong>of</strong> a plant is<br />

general <strong>and</strong> valid for any thermoeconomic formulation that uses equations<br />

connecting inlet <strong>and</strong> outlet flows <strong>of</strong> each component.<br />

Just as k * was defined as a marginal cost when production is modified, we can also<br />

obtain the marginal cost when the internal variables x are modified. Similarly,<br />

applying the chain rule, we get:<br />

∂B<br />

∂x<br />

0<br />

i<br />

=<br />

m<br />

∑<br />

j=<br />

1<br />

k<br />

*<br />

j<br />

∂g<br />

∂x<br />

j<br />

i<br />

(6.26)<br />

This equation expresses the effect on additional resource consumption when an<br />

internal parameter x i is modified <strong>and</strong> is the basis for the thermoeconomic diagnosis<br />

(explained in detail below).<br />

To determine the physical model <strong>of</strong> the system, a set <strong>of</strong> equations must be defined<br />

which relate the internal <strong>and</strong> external variables to the thermodynamic laws: mass,<br />

energy <strong>and</strong> entropy balances.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 139


<strong>Thermoeconomic</strong>s<br />

The most developed thermoeconomic optimization methodologies (Frangopoulos,<br />

1987, 1990; Von Spakovsky et al., 1993), use the Lagrange multipliers optimization<br />

method to calculate the marginal costs defined in the previous section. It can easily be<br />

proved (Serra, 1994; Reini, 1994) that the Lagrange multipliers are the marginal costs<br />

defined in equation (6.2), i.e:<br />

i = 1,..., m (6.27)<br />

This multiplier represents the variation <strong>of</strong> the objective function B 0 concerning the<br />

state variable B i.<br />

6.2.1 Marginal <strong>and</strong> average thermoeconomic costs<br />

Now, we will show that the marginal <strong>and</strong> average costs coincide when the<br />

characteristic equations <strong>of</strong> the system are first grade homogeneous functions<br />

concerning the extensive magnitude B. This is a very important result since the<br />

marginal <strong>and</strong> average costs can be calculated using the same procedure. This unifies<br />

accounting <strong>and</strong> optimization theories in a common mathematical formulation. The<br />

most important advantage is that variables <strong>and</strong> costs with different conceptual<br />

significance can be compared <strong>and</strong> better understood. Thus, the Exergetic Cost Theory<br />

(Valero, Lozano <strong>and</strong> Muñoz, 1986a), a cost accounting methodology which provides<br />

average costs, <strong>and</strong> <strong>Thermoeconomic</strong> Functional Analysis (Frangopoulos, 1983,<br />

1987), an optimization methodology which provides marginal costs, are particular<br />

cases <strong>of</strong> the Structural Theory. As a result <strong>of</strong> the integration <strong>of</strong> different approaches,<br />

some useful thermoeconomic applications have been developed, e. g. diagnosis<br />

operation <strong>and</strong> thermoeconomic optimization using the same mathematical formalism.<br />

As an illustration, consider a generic component or subsystem with several inlet <strong>and</strong><br />

outlet flows. For the sake <strong>of</strong> simplicity we will use a general subsystem with two inlet<br />

flows <strong>and</strong> two outlet flows (figure 6.3).<br />

FIGURE 6.3 Generic component scheme.<br />

B 1<br />

B 2<br />

λ i<br />

B<br />

=<br />

B<br />

∂<br />

∂<br />

0<br />

i<br />

4<br />

B1 3<br />

B1 4<br />

B2 140 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

3<br />

B2 B 3<br />

B 4


Calculating thermoeconomic costs<br />

The characteristic equations that describe component behavior are:<br />

B 1 = k 13 B 3 + κ 14 B 4<br />

B 2 = k 23 B 3 + κ 24 B 4<br />

(6.28)<br />

(6.29)<br />

These equations provide the amount <strong>of</strong> inlet resources (B 1, B 2) consumed to obtain<br />

each one <strong>of</strong> the outlet flows (B 3, B 4). This idea is easily understood if the component<br />

is made up <strong>of</strong> two subsystems. The equations modeling each subsystem are:<br />

B 13 = κ 13 B 3<br />

B 14 = κ 14 B 4<br />

B 23 = κ 23 B 3<br />

B 24 = κ 24 B 4<br />

(6.30a)<br />

(6.30b)<br />

(6.31a)<br />

(6.31b)<br />

Equations (6.30a, 6.31a) represent the resources needed to produce B 3 <strong>and</strong> Equations<br />

(6.30b, 6.31b) are the resources consumed to produce B 4. The total amount <strong>of</strong><br />

resources required to obtain B 3 is thus:<br />

<strong>and</strong> to obtain B 4:<br />

B 13 + B 23= κ 13 B 3 + κ 23 B 3<br />

B 14 + B 24 = κ 14 B 4 + κ 24 B 4<br />

According to Equation (6.1) the average cost <strong>of</strong> the outlet flows are:<br />

* κ13 B3 + κ23 B3 = ------------------------------------- = κ13 + κ23 k 3<br />

k 4<br />

B 3<br />

* κ14 B4 + κ24 B4 = ------------------------------------- = κ14 + κ24 B 4<br />

And the marginal cost <strong>of</strong> the outlet flows are:<br />

k<br />

k<br />

g<br />

B k 1 = ∂<br />

∂<br />

g<br />

+<br />

B ∂<br />

∂<br />

k = k + k<br />

* * 2 *<br />

3<br />

1<br />

2 13 23<br />

3<br />

3<br />

g<br />

B k 1 = ∂<br />

∂<br />

g<br />

+<br />

B ∂<br />

∂<br />

k = k +<br />

k<br />

* * 2 *<br />

4<br />

1<br />

2 14 24<br />

4<br />

4<br />

(6.32)<br />

(6.33)<br />

(6.34a)<br />

(6.34b)<br />

(6.35a)<br />

(6.35b)<br />

considering that the value <strong>of</strong> the marginal cost <strong>of</strong> the input flows (B 1, B 2) is equal to<br />

one. Since equations (6.34) are the same as equations (6.35), the average <strong>and</strong><br />

marginal costs <strong>of</strong> B 3 <strong>and</strong> B 4 coincide. Both kinds <strong>of</strong> costs coincide because the<br />

equations modeling the component are homogeneous functions <strong>of</strong> first order<br />

concerning the extensive magnitudes characterizing the outlet flows.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 141


<strong>Thermoeconomic</strong>s<br />

In this pro<strong>of</strong>, the cost <strong>of</strong> the inlet flows was unity. This is equivalent to considering<br />

that the subsystem was at the beginning <strong>of</strong> the productive process. The general<br />

mathematical formulation <strong>of</strong> the cost generated in a component is the same for each<br />

one <strong>and</strong> is not dependent on the position in the productive process. Thus, the results<br />

obtained are general.<br />

The average <strong>and</strong> marginal costs coincide because the equations modeling the<br />

components are first grade homogeneous functions concerning the extensive<br />

magnitude characterizing the outlet flows. The mass is the property determining<br />

whether a magnitude is extensive or not. If all equations modeling a system are first<br />

grade homogeneous functions concerning the mass, a simple substitution can<br />

transform those equations in homogeneous functions with respect to any extensive<br />

property. Thus, the marginal <strong>and</strong> average costs coincide if all equations modeling the<br />

behavior <strong>of</strong> the system are first grade homogeneous functions concerning the mass<br />

flow rate.<br />

6.2.2 Economic resources <strong>and</strong> thermoeconomic costs<br />

<strong>Thermoeconomic</strong> cost calculation considering the component capital cost Z, is<br />

similar to the above method but should be explained in more detail. The capital cost<br />

<strong>of</strong> each component Z can be considered an external flow <strong>of</strong> plant resources from the<br />

environment to the component (see figure 6.4). This will represent the monetary units<br />

per second needed to compensate the depreciation, maintenance cost <strong>and</strong> so on, <strong>of</strong> the<br />

component.<br />

FIGURE 6.4 Economic resources scheme.<br />

B 0<br />

Economic resources<br />

B i<br />

Z 1 = Z 1 (B 1 , B j , B h )<br />

According to marginal cost <strong>analysis</strong>, Z represents an environmental resource <strong>and</strong> can<br />

be h<strong>and</strong>led in the same mathematical way as energy resources. The amount <strong>of</strong><br />

resources consumed when manufacturing a device are, in fact, resources consumed to<br />

obtain the plant products. Some authors (Brodyansky et al., 1993; Le G<strong>of</strong>f, 1979)<br />

have developed methodologies to evaluate the total amount <strong>of</strong> resources consumed<br />

when building a process unit. Then the marginal unit cost ∂Z/∂B, can be considered a<br />

marginal consumption κ Zj.<br />

142 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

x 1<br />

B j<br />

B h


<strong>Thermoeconomic</strong> applications to thermoeconomic operation diagnosis <strong>and</strong> the optimization <strong>of</strong><br />

For the component depicted in figure 6.4 the characteristic equations are:<br />

B i = f (B j, κ ij) (6.36a)<br />

Z j = Z (B j, κ Zj) (6.36b)<br />

And the cost <strong>of</strong> the product is:<br />

k<br />

B<br />

B k i = ∂<br />

∂<br />

Z<br />

+<br />

B<br />

∂<br />

∂<br />

= k κ + κ<br />

* * j *<br />

j<br />

i<br />

i ij Zj<br />

j<br />

j<br />

If Z j is proportional to the production <strong>of</strong> the unit, or in other words its characteristic<br />

function is first order homogeneous, the marginal cost is equal to the average cost.<br />

But, unfortunately Z j is a non-linear function <strong>of</strong> the production in most cases.<br />

6.3 <strong>Thermoeconomic</strong> applications to thermoeconomic<br />

operation diagnosis <strong>and</strong> the optimization <strong>of</strong> complex<br />

energy systems<br />

Having defined the tools needed for a thermoeconomic <strong>analysis</strong> <strong>of</strong> a complex system,<br />

some applications to thermoeconomic diagnosis <strong>and</strong> optimization can be presented.<br />

The methodology is presented together with a simple application.<br />

6.3.1 Operation thermoeconomic diagnosis<br />

Diagnosis is the art <strong>of</strong> discovering <strong>and</strong> underst<strong>and</strong>ing signs <strong>of</strong> malfunction <strong>and</strong><br />

quantifying their effects. In the case <strong>of</strong> <strong>Thermoeconomic</strong>s, the effect <strong>of</strong> a malfunction<br />

is quantified in terms <strong>of</strong> additional resources consumed to obtain the same<br />

production, both in quality <strong>and</strong> in quantity.<br />

The main problem in energy system diagnosis can be summarized in the following<br />

question: Where, how <strong>and</strong> which part <strong>of</strong> the consumed resources can be saved by<br />

keeping the quantity <strong>and</strong> quality <strong>of</strong> the final products constant? To answer these<br />

questions, we need:<br />

• Procedures that accurately determine the state <strong>of</strong> the plant.<br />

• A theory to provide the concepts <strong>and</strong> tools to underst<strong>and</strong> <strong>and</strong> explain the causes<br />

<strong>of</strong> this state.<br />

The methodology presented in this paper applies Structural Theory to provide the<br />

tools to investigate the causes <strong>of</strong> the irreversibilities <strong>and</strong> the cost formation process.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 143


<strong>Thermoeconomic</strong>s<br />

In order to clarify the explanation <strong>of</strong> the proposed method we use a simple example (a<br />

more complex one can be found in Lerch, Royo <strong>and</strong> Serra, 1999), the co-generation<br />

plant depicted in figure 6.1, whose design <strong>and</strong> operational exergy flow values are<br />

shown in table 6.4. The plant has a co-generation gas turbine cycle <strong>and</strong> uses the<br />

turbine outlet gases as thermal energy in a heat recovery steam generator that<br />

produces steam (flow #7) together with the electric energy produced in the turbogenerator<br />

(flow #6).<br />

TABLE 6.4 Design <strong>and</strong> operation exergy flow values <strong>of</strong> the cogeneration plant (figure 6.1).<br />

Flow (kW) 1 2 3 4 5 6 7 8<br />

Design 11781 2704 9614 3831 2977 2500 2355 388<br />

Operation 11914 2758 9753 3887 3056 2500 2355 424<br />

6.3.1.1 Technical exergy saving<br />

Once the exergy flows have been supplied by an appropriate performance test or a<br />

model simulator, the irreversibilities in each productive unit can be obtained from the<br />

exergy balance. But not all exergy losses can be saved in practice. In fact, the<br />

potential exergy saving is limited by technical <strong>and</strong>/or economic constraints. It also<br />

depends on the decision level that limits the actions to be undertaken. In contrast to<br />

conventional thermodynamic <strong>analysis</strong>, <strong>Thermoeconomic</strong>s assumes a reference<br />

situation <strong>of</strong> the plant operating under design conditions. From this perspective, in the<br />

plant <strong>of</strong> figure 6.1, we see that only 133 kW, <strong>of</strong> the 7.06 MW <strong>of</strong> total irreversibilities<br />

can be saved with respect to design conditions.<br />

Therefore, the additional fuel consumption can be expressed as the difference<br />

between the resource consumption <strong>of</strong> the operating plant <strong>and</strong> the resource<br />

consumption for a reference or design condition with the same production objectives:<br />

∆FT = FT −FT 0<br />

<strong>and</strong> it can be broken up into the sum <strong>of</strong> the irreversibilities <strong>of</strong> each component:<br />

n<br />

∆FT = ∆IT = ⎛Ij<br />

−I<br />

⎞<br />

⎝ j⎠<br />

∆I<br />

=<br />

0<br />

∑ ∑<br />

j=<br />

1 j=<br />

1<br />

144 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

n<br />

(6.37a)<br />

(6.37b)<br />

However, even though the methods based on Second Law Analysis (Kotas, 1985) <strong>and</strong><br />

Technical Exergy Saving are useful to quantify the additional fuel consumption, they<br />

fail when trying to identify the real causes <strong>of</strong> the additional resources consumption.<br />

j


<strong>Thermoeconomic</strong> applications to thermoeconomic operation diagnosis <strong>and</strong> the optimization <strong>of</strong><br />

6.3.1.2 Impact on resources consumption<br />

The Fuel/Product diagram <strong>of</strong> the cogeneration plant is shown in figure 6.2. This<br />

diagram can be simplified by merging junction 1 <strong>and</strong> branching point 1 in a new<br />

fictitious component called junction–branching point (see figure 6.5). This new<br />

productive structure is slightly different than figure 6.2, <strong>and</strong> is more compact.<br />

The characteristic equations <strong>of</strong> this new productive structure are obtained as in the<br />

previous section applying equation (6.15)<br />

P = Ps + KP P<br />

FIGURE 6.5 Fuel / Product diagram <strong>and</strong> fuel <strong>and</strong> product exergy flows (kW) in design conditions for the cogeneration<br />

plant shown in figure 6.1.<br />

1<br />

1<br />

2<br />

2<br />

3-2<br />

F 0 F 1 F 2 F 3 F 4 Total<br />

P 0 0 11781 0 0 0 11781<br />

P 1 0 0 0 4156 2474 6631<br />

P 2 0 0 0 1627 968 2595<br />

P 3 2500 0 2977 0 0 5477<br />

P 4 2355 0 0 0 0 2355<br />

Total 4855 11781 2977 5783 3443<br />

8<br />

For the sake <strong>of</strong> simplicity we did not consider thermal <strong>and</strong> mechanical exergies as<br />

separate entities. Two auxiliary variables also appear r 1 = (B 3 – B 2)/B 3 <strong>and</strong> r 2 = B 3/B 2,<br />

which correspond to the part <strong>of</strong> the fuel <strong>of</strong> the turbine <strong>and</strong> HRSG coming from the<br />

combustor <strong>and</strong> the compressor respectively. Flow #8, produced in part in the<br />

combustor <strong>and</strong> in the compressor, also leaves the system as a residue. Only a part <strong>of</strong><br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 145<br />

5<br />

3-4<br />

4-8<br />

3<br />

4<br />

6<br />

7


<strong>Thermoeconomic</strong>s<br />

the entering gases to the turbine: B 3 – B 8 are used as a fuel <strong>of</strong> other components <strong>of</strong> the<br />

system. Therefore, only a part <strong>of</strong> the combustor’s <strong>and</strong> compressor’s product is used as<br />

a fuel for other components (useful product). Accordingly, figure 6.5 shows the chosen<br />

disaggregation scheme <strong>of</strong> the system <strong>and</strong> the Fuel/Product values for the design<br />

conditions.<br />

TABLE 6.5 Fuel/Product definition corresponding to figure 6.5<br />

No. Component Fuel Product Residue<br />

1 Combustor B 1 B 3 – B 2<br />

2 Compressor B 5 B 2<br />

3 Turbine B 3 – B 4 B 6<br />

4 HRSG B 4 – B 8 B 7 B 8<br />

In order to bring together the problem <strong>of</strong> the impact <strong>of</strong> resources consumption with<br />

thermoeconomic diagnosis we need to know the increase <strong>of</strong> the unit exergy<br />

consumption <strong>of</strong> each component <strong>of</strong> the plant. A performance test or a simulator<br />

provides the real values <strong>of</strong> the unit consumptions which are then compared with the<br />

design values.<br />

TABLE 6.6 Increase <strong>of</strong> unit consumption. (100 ∆κ ij ).<br />

∆κ e 0.4006 0.0000 0.0000 0.0000<br />

∆ KP<br />

0.0000 0.0000 –0.1667 0.3857<br />

0.0000 0.0000 0.1593 0.4636<br />

0.0000 1.1147 0.0000 0.0000<br />

0.0000 0.0000 0.0000 0.0000<br />

∆k 0.4006 1.1147 -0.0074 0.8493<br />

The values <strong>of</strong> the unit exergetic consumption increase are found as: ∆κ ij = κ ij (x) – κ ij<br />

(x 0 ). Table 6.6 shows the ∆κ ij values for the plant in figure 6.1.<br />

Equation (6.19) is used to obtain the increment <strong>of</strong> the total resources <strong>of</strong> an operating<br />

plant regarding the reference conditions:<br />

t 0<br />

t<br />

∆ = ∆ κ P + κ ∆P<br />

F T<br />

e<br />

146 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

e<br />

(6.38)


<strong>Thermoeconomic</strong> applications to thermoeconomic operation diagnosis <strong>and</strong> the optimization <strong>of</strong><br />

The increase <strong>of</strong> the component production from equation (6.16) may be expressed in<br />

terms <strong>of</strong> the unit exergy consumption as:<br />

hence, applying equation (6.17), we obtain:<br />

(6.39)<br />

(6.40)<br />

If we want to analyze the fuel impact due to an increment <strong>of</strong> the exergy unit<br />

consumption <strong>of</strong> the components, equation (6.38) could be written as:<br />

If no change in the total production <strong>of</strong> the plant is assumed, then:<br />

or in scalar format:<br />

0<br />

∆P = ∆P + ∆ KP P + KP ∆P<br />

s<br />

∆P | P〉<br />

∆Ps ∆ 〈 KP〉<br />

P 0<br />

= ⎛ + ⎞<br />

⎝ ⎠<br />

t 0 t * 0 t *<br />

∆ = ∆ κ P + κ ∆ KP P + κ ∆P<br />

F T<br />

e<br />

⎛ t t ⎞<br />

∆FT= ⎜∆<br />

κe+ κ P ∆ ⎟<br />

⎝<br />

⎠<br />

* KP P 0<br />

n ⎛ n ⎞<br />

*<br />

∆FT = ∑ ⎜ k j ∆ ji Pi<br />

⎜∑<br />

P, κ ⎟<br />

i = 1 ⎝ j=<br />

0 ⎠<br />

P<br />

0<br />

P s<br />

(6.41)<br />

(6.42a)<br />

(6.42b)<br />

Using the above equation, the additional resource consumption ∆F T (also called Fuel<br />

Impact; Reini, 1994) can be expressed as the sum <strong>of</strong> the contributions <strong>of</strong> each<br />

component.<br />

The variation <strong>of</strong> the exergetic unit consumption <strong>of</strong> each component increases its<br />

resources consumption <strong>and</strong> its irreversibilities in a quantity ∆κ ji Pi , which we call,<br />

malfunction. Consequently, this implies an additional consumption <strong>of</strong> external<br />

resources given by , which is also named the malfunction cost.<br />

Therefore, the total fuel impact can be written as the sum <strong>of</strong> the fuel impact or<br />

malfunction cost <strong>of</strong> each component, as shown in equation (6.42b).<br />

0<br />

* 0<br />

kP, j ∆κ ji Pi<br />

The proposed method provides the exact values <strong>of</strong> the additional resource<br />

consumption <strong>of</strong> each component malfunction for any operational state. Other<br />

methods, such as the Theory <strong>of</strong> Perturbations (Lozano et al., 1996), only provide an<br />

approximate predictive value, based on marginal costs (Lagrange multipliers) which<br />

is valid for an operating state close to the reference conditions.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 147


<strong>Thermoeconomic</strong>s<br />

FIGURE 6.6 Fuel impact <strong>and</strong> technical saving.<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Figure 6.6 compares the fuel impact <strong>and</strong> the increase <strong>of</strong> irreversibilities or the<br />

technical exergy saving <strong>of</strong> each component <strong>and</strong> also compares (first column) the<br />

malfunction <strong>and</strong> the fuel impact for each component. Three malfunctions in the plant<br />

are shown in the combustor, the compressor <strong>and</strong> the HRSG. The largest<br />

irreversibilities increase is in the combustor, but the largest fuel impact is in the<br />

compressor. The question that arises is: What causes the irreversibilities increase <strong>and</strong><br />

the fuel impact, <strong>and</strong> how are they related?<br />

6.3.1.3 Malfunction <strong>and</strong> dysfunction <strong>analysis</strong><br />

We have shown that there is no direct relationship between the increase <strong>of</strong> the<br />

irreversibilities <strong>and</strong> fuel impact. The more advanced the production process is, the<br />

greater the cost <strong>of</strong> the irreversibility malfunction <strong>and</strong>, as a consequence, the greater<br />

its fuel impact.<br />

Furthermore, the degradation <strong>of</strong> a component will force other components to adapt<br />

their behavior in order to maintain their production conditions <strong>and</strong> modify their<br />

irreversibilities. Figure 6.7 shows how an increase <strong>of</strong> the unit consumption <strong>of</strong> a<br />

component will not only increase the irreversibilities on it but also the irreversibilities<br />

<strong>of</strong> the previous component.<br />

FIGURE 6.7 Malfunction <strong>and</strong> fuel impact.<br />

Combustor Compressor Turbine HRSG<br />

∆F 1<br />

F 1<br />

∆I 1<br />

I 1<br />

∆P 1<br />

P 1<br />

Fuel Impact<br />

Malfunction<br />

Technical Saving<br />

148 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

∆F 2<br />

F 2<br />

∆I 2<br />

I 2<br />

P 2


<strong>Thermoeconomic</strong> applications to thermoeconomic operation diagnosis <strong>and</strong> the optimization <strong>of</strong><br />

The irreversibility increase <strong>of</strong> a generic system’s component is given by:<br />

∆I = ∆K D P 0 + (K D – U D) ∆P (6.43)<br />

From the above expression, we can distinguish two types <strong>of</strong> irreversibilities:<br />

Endogenous irreversibility or malfunction produced by an increase <strong>of</strong> the unit<br />

consumption <strong>of</strong> the component itself:<br />

Exogenous irreversibility or dysfunction induced in the component by the<br />

malfunction <strong>of</strong> other subsystems, which forces it to consume more local resources to<br />

obtain the additional production required by the other components:<br />

The malfunction only affects the behavior <strong>of</strong> the components; the dysfunction is a<br />

result <strong>of</strong> how the components adapt themselves to maintain the total production.<br />

Now we will consider the causes <strong>and</strong> effects <strong>of</strong> the irreversibilities systems <strong>and</strong><br />

introduce a new method to compute the fuel impact <strong>of</strong> a malfunction <strong>and</strong> its effect. In<br />

other words, how to compute the dysfunction on the rest <strong>of</strong> the system components.<br />

If we substitute ∆P from equation (6.40) then the irreversibility increase <strong>of</strong> each<br />

component, equation (6.43) is written in terms <strong>of</strong> the unit consumption as:<br />

or in scalar format:<br />

0 0<br />

MF = P ∆k= P ∆κ i<br />

j=<br />

i i i ji<br />

0<br />

DF = ( k −1)<br />

∆P<br />

i i i<br />

n<br />

∑<br />

⎛<br />

⎞<br />

∆I= ⎜∆KD+<br />

I ∆ KP ⎟P<br />

⎝<br />

⎠<br />

n<br />

∑ ∑<br />

∆I = P ∆κ + φ ∆κ<br />

P<br />

i i ji<br />

j=<br />

1<br />

jh , = 1<br />

n<br />

0<br />

ih hj<br />

0<br />

j<br />

(6.44)<br />

i = 1,…, n (6.45)<br />

The first part <strong>of</strong> the previous expression corresponds to the component malfunction,<br />

<strong>and</strong> the last part to the dysfunction. If we denote:<br />

n<br />

DFij ∑ φih ∆κhj<br />

Pj<br />

=<br />

h=<br />

1<br />

0<br />

(6.46)<br />

DFij represents the part <strong>of</strong> i–th component dysfunction generated by component –j–,<br />

where φih are the coefficients <strong>of</strong> the irreversibility matrix operator | I〉<br />

for the actual<br />

operation values. The above expression shows how a malfunction Pj ∆κhj, on the j-th<br />

component, generates a dysfunction on the i–th component proportional to the φih coefficients, which represent the weight <strong>of</strong> the malfunction effect. The coefficient φih <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 149


<strong>Thermoeconomic</strong>s<br />

does not depend on the malfunction amount, but only on the unit consumption <strong>of</strong> the<br />

components in the operating state. Therefore, the dysfunction cannot be corrected by<br />

itself but decreases the malfunction which generated it.<br />

The technical exergy saving <strong>of</strong> component –i–, equation (6.45) can be written as the<br />

sum <strong>of</strong> its malfunction <strong>and</strong> the dysfunction generated by other components <strong>of</strong> the<br />

system:<br />

i i ∑ ij<br />

1<br />

∆I = MF + DF<br />

j=<br />

i = 1,…, n (6.47)<br />

The graph in figure 6.8 describes the cause <strong>of</strong> the irreversibilities increase in the gas<br />

turbine cycle (<strong>of</strong> the example) as the sum <strong>of</strong> the malfunctions <strong>and</strong> the dysfunction<br />

generated by the rest <strong>of</strong> the components<br />

FIGURE 6.8 Analysis <strong>of</strong> the irreversibility causes (kW).<br />

80<br />

60<br />

40<br />

20<br />

0<br />

∆ I1<br />

Fuel impact <strong>and</strong> dysfunction<br />

n<br />

∆ I2<br />

For a specified constant quality <strong>and</strong> quantity <strong>of</strong> total production, the fuel impact<br />

(6.42b) could be written as the sum <strong>of</strong> the malfunctions <strong>and</strong> dysfunctions <strong>of</strong> all the<br />

plant components:<br />

150 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(6.48)<br />

If we rearrange the previous expression, grouping by component production, we<br />

obtain:<br />

∆ I3<br />

n<br />

n ⎛<br />

n ⎞<br />

∆FT = ∑ ∆Ii<br />

= ∑ ⎜MFi<br />

+ DFij<br />

⎜ ∑ ⎟<br />

i=<br />

1 i=<br />

1<br />

⎝ j=<br />

1 ⎠<br />

HRSG<br />

Turbine<br />

Compressor<br />

Combustor<br />

Malfunction<br />

∆ I4


<strong>Thermoeconomic</strong> applications to thermoeconomic operation diagnosis <strong>and</strong> the optimization <strong>of</strong><br />

n ⎛<br />

n ⎞<br />

∆FT= ∑ ⎜∆ki+<br />

jh ∆ hi Pi<br />

⎜ ∑φ<br />

κ ⎟<br />

i=<br />

1 ⎝ jh , = 1 ⎠<br />

(6.49)<br />

Therefore, the fuel impact or the malfunction cost <strong>of</strong> each component is given by the<br />

sum <strong>of</strong> the malfunction <strong>and</strong> the dysfunction:<br />

i i ∑ hi<br />

h=<br />

1<br />

*<br />

MF = MF + DF<br />

i = 1,…, n (6.50)<br />

If we compare the previous equations with the fuel impact equation (6.42b), we find a<br />

relationship between the unit cost <strong>of</strong> production <strong>and</strong> the irreversibility dysfunction<br />

coefficients, given by:<br />

n<br />

*<br />

k Pj , = 1 + ∑ φij<br />

i=<br />

1<br />

j = 1,…, n (6.51)<br />

The above expression is an alternative method to compute the unit cost <strong>of</strong> the product<br />

as the sum <strong>of</strong> the contribution <strong>of</strong> the irreversibilities <strong>of</strong> each component. Table 6.7<br />

shows the irreversibility matrix operator coefficients <strong>and</strong> unit cost <strong>of</strong> the component<br />

product for an operating gas turbine plant.<br />

TABLE 6.7 Irreversibility matrix <strong>and</strong> unit cost <strong>of</strong> product.<br />

| I〉<br />

k P *<br />

n<br />

0.7807 1.0469 0.9037 1.2586<br />

0.0000 0.2422 0.0723 0.1007<br />

0.0000 0.0988 0.0853 0.0411<br />

0.0000 0.0000 0.0000 0.4704<br />

1.7807 2.3880 2.0614 2.8708<br />

A graph <strong>of</strong> the fuel impact for each component is shown in figure 6.9. Note that the<br />

dysfunction becomes even greater than its own malfunction as the production process<br />

proceeds. The cost <strong>of</strong> the malfunction in the compressor <strong>and</strong> HRSG includes the<br />

dysfunction generated, for the most part, in the combustor.<br />

The sum <strong>of</strong> the dysfunctions generated by a component:<br />

n<br />

i = ∑ ji<br />

j=<br />

1<br />

DI DF<br />

n<br />

i = 1,…, n (6.52a)<br />

could be written as: DIi = ⎛ *<br />

k P j − ⎞ 0<br />

∑ ⎝ , 1<br />

⎠<br />

∆κ<br />

ji Pi<br />

(6.52b)<br />

j=<br />

1<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 151<br />

0


<strong>Thermoeconomic</strong>s<br />

FIGURE 6.9 Analysis <strong>of</strong> fuel impact (kW).<br />

80<br />

60<br />

40<br />

20<br />

0<br />

∆ I4<br />

∆ I3<br />

∆ I2<br />

∆ I1<br />

MF<br />

Combustor Compressor Turbine HRSG<br />

Therefore, the dysfunction generated by a component (as with the fuel impact)<br />

depends on the malfunction <strong>and</strong> the position <strong>of</strong> the component in the productive<br />

process, which is, in turn, characterized by the unit cost <strong>of</strong> the resources required by<br />

the component.<br />

The relationship between irreversibility increase <strong>and</strong> fuel impact can be represented<br />

by a double input table (see table 6.8). The dysfunction table containing the DF ij<br />

elements can be computed in a compact matrix form using the expression:<br />

[ DF] I KP P<br />

= ∆ 0<br />

D<br />

TABLE 6.8 Malfunction <strong>and</strong> dysfunction table in (kW).<br />

Combustor Compressor Turbine HRSG DF Malfunction Total<br />

∆I 1 0.000 26.140 2.004 18.520 46.664 26.562 73.226<br />

∆I 2 0.000 2.092 2.113 2.644 6.849 28.925 35.774<br />

∆I 3 0.000 2.467 0.862 1.079 4.408 –0.408 4.000<br />

∆I 4 0.000 0.000 0.000 0.000 0.000 20.000 20.000<br />

DI 0.000 30.699 4.979 22.243 57.921<br />

Malfunction 26.562 28.925 –0.408 20.000 75.079<br />

Total 26.562 59.624 4.571 42.243 133.000<br />

152 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> applications to thermoeconomic operation diagnosis <strong>and</strong> the optimization <strong>of</strong><br />

Each cell represents the DF ij dysfunction. The DI column represents the sum <strong>of</strong> the<br />

dysfunctions generated by each component, <strong>and</strong> the DF row is the sum <strong>of</strong> the<br />

dysfunctions generated in each component. The total sum by columns represents the<br />

Fuel Impact <strong>of</strong> each component, <strong>and</strong> the total sum by rows is the irreversibility<br />

increase. The methodology proposed in this section is summarized in the table<br />

mentioned above. It is a powerful tool to find the causes <strong>and</strong> effects <strong>of</strong> variations from<br />

the design conditions <strong>of</strong> a plant <strong>and</strong> to study, classify <strong>and</strong> assign the role <strong>of</strong> each<br />

system unit.<br />

6.3.1.4 Intrinsic <strong>and</strong> induced malfunctions<br />

Using the above method we can identify <strong>and</strong> quantify malfunction effects. For<br />

example, we found three malfunctions in the gas turbine cycle (figure 6.1): one each<br />

in the combustor, compressor <strong>and</strong> HRSG. But, What are the causes <strong>of</strong> the<br />

malfunctions? In fact, the actual operation values shown in table 6.4 correspond to a<br />

1% decrease in compressor isoentropic efficiency. This means that HRSG <strong>and</strong><br />

combustor efficiencies can be changed by varying compressor efficiency.<br />

How do we approach this problem? The relationship between operation <strong>and</strong><br />

efficiency <strong>of</strong> the components could be analyzed using a simulator. If all the pant<br />

components were isolated, the efficiencies <strong>of</strong> those components would be<br />

independent variables (Lozano et al., 1996). So we will assume that there is an<br />

operating parameter x r affecting the efficiency <strong>of</strong> the i-th component <strong>of</strong> the plant <strong>and</strong><br />

thus, in most cases, also indirectly affecting the efficiencies <strong>of</strong> the other plant process<br />

units.<br />

Once the relationship between unit exergy consumption <strong>and</strong> the operating parameters<br />

is known, the above methodology can be applied to distinguish the effect <strong>of</strong> an<br />

operating parameter on the internal economy <strong>of</strong> a component, i.e. its malfunction <strong>and</strong><br />

the cost <strong>of</strong> its malfunction.<br />

Plant operating parameters could be classified according to their effect on the<br />

efficiency <strong>of</strong> the components <strong>of</strong> the system:<br />

Local variables: They mainly affect the behavior <strong>of</strong> the component related to the<br />

variable, e.g, the isoentropic efficiency <strong>of</strong> a turbine. From a practical point <strong>of</strong> view, a<br />

variable is considered local <strong>and</strong> therefore related to a subsystem. The total fuel impact<br />

due to its perturbation is basically located in this component.<br />

Global <strong>and</strong>/or zonal variables: This is the case when an operating parameter cannot<br />

be associated with a specific component. We must identify them as operating set<br />

points, environmental parameters <strong>and</strong> the production load or fuel quality.<br />

In this thesis we will focus our <strong>analysis</strong> on local variables <strong>and</strong> how they affect<br />

additional fuel consumption <strong>and</strong> the other plant components. This <strong>analysis</strong> is, in fact,<br />

the next step in the thermoeconomic diagnosis.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 153


<strong>Thermoeconomic</strong>s<br />

Unfortunately the problem <strong>of</strong> locating causality <strong>of</strong> losses in a structure is rather more<br />

complex than locating malfunctions <strong>and</strong> dysfunctions.<br />

When a plant unit deteriorates (when its behavior is degraded) its physical variables<br />

are modified, its efficiency is decreased <strong>and</strong> its unit exergy consumption increases.<br />

The unit exergy consumption increase <strong>of</strong> each component, due to the variation <strong>of</strong> an<br />

operating parameter x r, is:<br />

r<br />

∆κ = κ ( x + ∆x)<br />

−κ<br />

( x )<br />

ij<br />

ij 0 r ij 0<br />

Therefore, it will be possible to approximate the malfunction <strong>of</strong> a component as the<br />

sum <strong>of</strong> the contributions <strong>of</strong> each operating parameter:<br />

r<br />

MFi ≅ ∑ ∑ ∆κ ji Pi<br />

r<br />

n<br />

j=<br />

1<br />

154 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

0<br />

(6.53)<br />

According to the classification <strong>of</strong> operating parameters, the intrinsic malfunction is<br />

that part <strong>of</strong> the component malfunction due to the degradation/improvement <strong>of</strong> the<br />

component itself, which is, in turn, due to variation <strong>of</strong> local operating parameters:<br />

L<br />

r<br />

MFi ≡ ∑ ∑ ∆κ ji Pi<br />

n<br />

r∈Lij= 1<br />

0<br />

(6.54)<br />

A system malfunction or improvement does not only have consequences upstream<br />

(by trying to see the variation in consumption <strong>of</strong> used resources) but also<br />

downstream. Clearly the degradation or improvement <strong>of</strong> a system’s flow entry<br />

conditions will affect its efficiency to a greater or lesser extent. This will modify the<br />

production <strong>and</strong> affect the next component.<br />

Not only are there dysfunctions when there is an intrinsic malfunction. There are also<br />

induced malfunctions, that can decisively affect the system's behavior. For example,<br />

using the throttle valve in a power plant can destroy a small additional amount <strong>of</strong><br />

exergy but the downstream effects on turbine efficiencies can be quite serious.<br />

Thus, the difference between total component malfunction <strong>and</strong> intrinsic malfunction<br />

is called induced malfunction. It is due to the degradation <strong>of</strong> other plant components<br />

which provoke a variation in the unit consumption <strong>of</strong> that component:<br />

G<br />

MF = MF −MF<br />

i<br />

L<br />

i i<br />

(6.55)<br />

This phenomenon is not foreseen in classic linear thermoeconomic theory. The<br />

average cost obtained from the most rigorous disaggregation <strong>analysis</strong> can never<br />

predict induced malfunctions <strong>and</strong> dysfunctions will only be predicted in cases where<br />

the hypothesis <strong>of</strong> linearity <strong>and</strong> continuity holds.


<strong>Thermoeconomic</strong> applications to thermoeconomic operation diagnosis <strong>and</strong> the optimization <strong>of</strong><br />

Malfunction matrix<br />

It is important to know the fuel impact associated with the variation <strong>of</strong> each physical<br />

parameter when a malfunction occurs.<br />

The fuel impact <strong>of</strong> an operating parameter on the whole plant can be calculated using<br />

the simulator but the latter does not provide information about the effects on other<br />

plant components. A deterioration in a component (intrinsic malfunction) can modify<br />

the efficiencies <strong>of</strong> other plant components.<br />

Information about interactions among different plant components can be obtained<br />

with the methodology presented here. It is basically contained in the so called<br />

malfunction matrix, or 〈∆KP〉 matrix. This matrix can relate any operating parameter<br />

with all the possible malfunctions. Note that the overall impact on resources (fuel<br />

impact) can be written as:<br />

* r 0<br />

* r<br />

∆F ≡ k ∆κ P + k ∆κ<br />

P<br />

T P, j ji i<br />

r∈ Li<br />

j=<br />

0<br />

r∉Li j=<br />

0<br />

(6.56)<br />

Where the first term is the fuel impact associated with the intrinsic malfunction <strong>and</strong><br />

the last term is the fuel impact associated with the induced malfunctions <strong>and</strong> ∆κ ij are<br />

elements <strong>of</strong> the ∆ 〈KP〉 matrix. The ∆ 〈KP〉 matrix has been built for each parameter<br />

(see Chapter 7) in a variational <strong>analysis</strong>.<br />

In a real power plant, the most general case is when several plant components suffer<br />

simultaneous efficiency deviations. The total fuel impact can be calculated from the<br />

∆ 〈KP〉 matrix associated with each physical parameter <strong>and</strong> its causes can be<br />

explained <strong>and</strong> quantified component by component.<br />

This operation is completely new in <strong>Thermoeconomic</strong>s or in any energy <strong>analysis</strong><br />

technique. Thus, the malfunction matrix has a very important engineering application<br />

<strong>and</strong> also introduces new theoretical ideas in <strong>Thermoeconomic</strong>s (see Chapter 7).<br />

6.3.2 <strong>Thermoeconomic</strong> optimization<br />

n<br />

∑ ∑ ∑ ∑<br />

n<br />

Pj , ji<br />

Here we describe strategies for optimizing complex systems as proposed by Lozano<br />

et al. (1996). They are based on sequential optimization from component to<br />

component using the <strong>Thermoeconomic</strong> Isolation Principle (Evans, 1980).<br />

A component <strong>of</strong> a thermal system is thermoeconomically isolated from the rest <strong>of</strong> the<br />

system if the product <strong>of</strong> the component <strong>and</strong> the unit cost <strong>of</strong> its resources (internal<br />

product <strong>and</strong>/or external resources) are constant <strong>and</strong> known quantities. If a unit <strong>of</strong> a<br />

thermal system is thermoeconomically isolated, the unit may be optimized by itself<br />

(without considering the modifications <strong>of</strong> other variables <strong>of</strong> the rest <strong>of</strong> the system)<br />

<strong>and</strong> the optimun solution obtained for the unit coincides with the optimum solution<br />

for the whole system.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 155<br />

0<br />

i


<strong>Thermoeconomic</strong>s<br />

Of course, TI (<strong>Thermoeconomic</strong> Isolation) is an ideal condition which cannot be<br />

achieved in most <strong>of</strong> the real systems: Pj <strong>and</strong> k * P,i change when design variables <strong>of</strong><br />

other components change, due to feedback. But the more constant Pj <strong>and</strong> k * P,i are, the<br />

closer to TI conditions <strong>and</strong> the fewer iteration loops needed to achieve the optimal<br />

solution for the whole system. Thus, the goal is not to achieve TI but to approach it as<br />

much as possible in order to obtain maximum advantages, which include:<br />

1. Improvements <strong>and</strong> optimal design <strong>of</strong> individual units in highly interdependent<br />

complex systems are greatly facilitated, as well as <strong>of</strong> whole systems.<br />

2. The designers can be specialized <strong>and</strong> their efforts concentrated on designing the<br />

variables <strong>of</strong> single units, while resting assured that these efforts yield optimum<br />

design <strong>and</strong>/or improve the overall system.<br />

3. The convergence <strong>of</strong> the solution is faster.<br />

To optimize individual units, the objective function <strong>of</strong> the cost <strong>of</strong> product <strong>of</strong> the<br />

component –j– could be defined as:<br />

Min<br />

κ<br />

n<br />

⎛ * ⎞<br />

⎜∑κij kPi , ⎟ Pj<br />

⎝ ⎠<br />

i = 0<br />

156 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(6.57)<br />

where the unit cost <strong>of</strong> the input resources kPi , <strong>and</strong> the production Pj are known <strong>and</strong><br />

constant.<br />

In real world optimization problems, the design free variables do not necessarily<br />

coincide with the technical production coefficients. In practice there will be a<br />

function <strong>of</strong> the actual design free variables which can be named –x–.<br />

We say that a free variable x is a local variable <strong>of</strong> a subsystem –j– when the<br />

production coefficients κ ij <strong>of</strong> this subsystem only depend on x. When a design<br />

variable is attached to several subsystems, the previous expression must be extended<br />

to all concerned subsystems.<br />

To determine whether a design free variable is local or not <strong>and</strong> which components are<br />

involved, the cost resource impact <strong>of</strong> the design variables to each component can be<br />

computed:<br />

x<br />

∆C0, j<br />

*<br />

kPi ,<br />

∂κ n<br />

ij<br />

∑ --------<br />

∂x<br />

i = 0<br />

∂κZP, j<br />

⎛ ⎞<br />

=<br />

⎜ + ------------- ⎟ P ∆x<br />

⎝ ∂x ⎠<br />

*<br />

(6.58)


<strong>Thermoeconomic</strong> applications to thermoeconomic operation diagnosis <strong>and</strong> the optimization <strong>of</strong><br />

<strong>and</strong> the ratio calculated:<br />

x<br />

εj ∆ C0, j<br />

=<br />

------------------------<br />

n<br />

∑<br />

i = 1<br />

(6.59)<br />

If this ratio is equal (or close) to 1, the design variable is local for component –j–, if it<br />

is equal (or close) to zero, the design variable is independent <strong>of</strong> the referred j<br />

component. In other cases the design variable involves several components.<br />

These ideas could be used to design a strategy for global optimization problems:<br />

1. Determine which variables are local <strong>and</strong> which are regional (involve several components)<br />

2. Determine a sequence for local optimization <strong>of</strong> each component<br />

3. Take an initial value <strong>of</strong> the design variables<br />

4. Calculate technical production coefficients <strong>and</strong> unit product cost<br />

5. Find optimum values for local variables<br />

x<br />

x<br />

∆ C0i ,<br />

6. Find optimum values for global variables<br />

7. Iterate from (3) to convergence when design variables or unit product cost do not<br />

vary in the next iteration. In each iteration the unit cost <strong>of</strong> total product must<br />

decrease.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 157


CHAPTER 7<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong><br />

a dual-purpose power <strong>and</strong><br />

desalination plant<br />

The basic concepts <strong>and</strong> fundamentals <strong>of</strong> <strong>Thermoeconomic</strong>s were explained in<br />

Chapter 6 <strong>and</strong> will now be applied to a dual-purpose power <strong>and</strong> desalination plant; the<br />

most important contribution <strong>of</strong> this Ph. D. Thesis. During the 60’s <strong>and</strong> early 70’s<br />

Evans (1962), Tribus (Tribus et al., 1960; Tribus <strong>and</strong> Evans, 1963) <strong>and</strong> El-Sayed (El-<br />

Sayed <strong>and</strong> Aplenc, 1970; El-Sayed <strong>and</strong> Evans, 1970) laid down the seminal ideas <strong>of</strong><br />

<strong>Thermoeconomic</strong>s <strong>and</strong> applied them to the desalination processes. Tribus first<br />

proposed the term ‘ <strong>Thermoeconomic</strong>s’.<br />

Since then, <strong>Thermoeconomic</strong> techniques have<br />

been developed <strong>and</strong> applied mostly to power plants. This thesis represents the most<br />

complete <strong>and</strong> rigorous thermoeconomic <strong>analysis</strong> ever made on a complex energy<br />

system <strong>and</strong> more specifically on a dual-purpose power <strong>and</strong> desalination plant. It<br />

encompasses the whole range between an energy audit based on a detailed cost<br />

<strong>analysis</strong>, up to a thermoeconomic optimization, via a thermoeconomic diagnosis <strong>of</strong><br />

several plant component failures.<br />

The first section <strong>of</strong> this chapter includes the resolution <strong>of</strong> the thermoeconomic model<br />

for the power <strong>and</strong> desalination plant. The steps to build <strong>and</strong> solve the thermoeconomic<br />

model are described in detail.<br />

The second section contains an in depth cost <strong>analysis</strong> <strong>of</strong> the most significant operating<br />

modes governing the power <strong>and</strong> desalination plant (including operational <strong>and</strong><br />

investment capital costs) in order to quantify the efficiency <strong>of</strong> each operation mode.<br />

This is used to calculate the physical (<strong>and</strong> therefore more realistic) value <strong>of</strong> the<br />

resources consumed to produce every flowstream inside the plant, which is the key to<br />

an energy audit. An inefficient process can be located <strong>and</strong> quantified in terms <strong>of</strong> fuel<br />

plant consumption. Eight different operating modes were considered in the dualpurpose<br />

plant, covering the whole range <strong>of</strong> the diverse combinations:<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


160<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

• In the first case, the plant only produced electricity. The second case was the<br />

opposite: the plant was like a pure distillation unit, producing only desalted<br />

water.<br />

• The third to sixth cases studied the effect <strong>of</strong> partial load operation on the<br />

efficiency <strong>of</strong> the installation, starting from maximum production to the minimum<br />

load <strong>of</strong> water <strong>and</strong> electricity dem<strong>and</strong>.<br />

• The seventh <strong>and</strong> eighth cases considered the effect <strong>of</strong> the cleaning ball system on<br />

the MSF evaporators. In both cases, some live steam was throttled in the HP<br />

reduction station corresponding to the maximum load <strong>of</strong> extracting live steam to<br />

a second MSF unit.<br />

Non-operating costs were added to the calculated exergy costs. We compared our<br />

thermoeconomic method with other indirect methods that calculate product costs as<br />

the accounting <strong>of</strong> expenses in plant exploitation: fuel, maintenance, amortization,<br />

etc., divided by the total plant production.<br />

The third section <strong>of</strong> this chapter describes a complete thermoeconomic diagnosis <strong>of</strong><br />

the inefficiencies in the units <strong>of</strong> the power <strong>and</strong> desalination plant. Not only was the<br />

additional fuel consumption due to the inefficiency calculated (impact on fuel<br />

<strong>analysis</strong>), but also the effect on the behavior <strong>of</strong> other plant units. This effect was<br />

separated in different contributions over the rest <strong>of</strong> devices: malfunctions (induced<br />

<strong>and</strong> intrinsic malfunctions) <strong>and</strong> dysfunctions. Four different loads in the power plant<br />

were considered <strong>and</strong> two distillate productions in the MSF plant. These examples<br />

encompass the most significant operating conditions. Each study considered five<br />

inefficiencies corresponding to five components <strong>of</strong> the power plant <strong>and</strong> three<br />

inefficiencies in the MSF plant.<br />

The fourth section applies the thermoeconomic technique based on local<br />

optimization. The local optimization <strong>of</strong> energy systems is very valuable to find the<br />

optimum operating conditions. The plant can be optimized by minimizing the cost <strong>of</strong><br />

the product <strong>of</strong> each unit, starting from real operating conditions.<br />

The fifth section analyzes the concepts <strong>of</strong> price <strong>and</strong> cost. They were distinguished in<br />

order to obtain the maximum benefit in plant exploitation.<br />

Finally, the last section contains the conclusions <strong>and</strong> some ‘ operation recommendations’<br />

from the thermoeconomic <strong>analysis</strong>. These are very useful to guide managers in<br />

saving energy <strong>and</strong> achieving a more cost-effective operation <strong>of</strong> a dual-purpose plant<br />

operation.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 7.1<br />

<strong>Thermoeconomic</strong> model<br />

7.1 <strong>Thermoeconomic</strong> model<br />

A thermoeconomic model mathematically represents the productive structure <strong>of</strong> a<br />

plant. This structure is a graphical representation <strong>of</strong> the resource distribution. Its<br />

flows describe the productive relationship among components based on the physical<br />

structure, although they are not forced to coincide with the existing physical flows <strong>of</strong><br />

the plant.<br />

The thermoeconomic model should logically be defined after the physical structure<br />

(section 7.1.2). Then the productive structure is built (section 7.1.3) along with the<br />

system <strong>of</strong> characteristic equations that mathematically describe the productive<br />

structure <strong>of</strong> the plant (section 7.1.4). Before considering the complex thermoeconomic<br />

model <strong>of</strong> the dual plant, a very simple thermoeconomic model <strong>of</strong> a cogeneration<br />

system is included in section 7.1.1. It is a simple example <strong>of</strong> how to build<br />

a thermoeconomic model <strong>and</strong> calculate the cost <strong>of</strong> live steam, water <strong>and</strong> electricity.<br />

These can be compared with other methodologies that only account for the cost <strong>of</strong> the<br />

final products (water <strong>and</strong> electricity) with external information or other<br />

simplifications (see section 7.2.5).<br />

7.1.1 A simple co-generation system<br />

A steam generator (boiler), a steam turbine <strong>and</strong> the MSF plant can represent a very<br />

simple dual-purpose desalination plant. Auxiliary non-producer elements like heaters,<br />

pumps or condensers are not included in the scheme. The productive structure in<br />

figure 7.1 can be built using the F-P definitions in table 7.1. The availability <strong>of</strong> the<br />

steam generated in the boiler is sent to the two productive units (steam turbine <strong>and</strong><br />

MSF desalination unit). The fuel <strong>and</strong> product definition <strong>and</strong> the characteristic <strong>and</strong><br />

exergy cost equations <strong>of</strong> every component <strong>of</strong> the system are included in table 7.1.<br />

Productive structure <strong>of</strong> the simple co-generation system.<br />

C 1<br />

1<br />

Boiler<br />

B 1<br />

A<br />

B 1 – B 2<br />

B 2<br />

2<br />

Steam turbine<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

3<br />

MSF unit<br />

W<br />

D<br />

161


TABLE 7.1<br />

TABLE 7.2<br />

162<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

Fuel, product, characteristic equation <strong>and</strong> exergy cost balance in the simple co-generation<br />

system.<br />

1 Boiler C<br />

1<br />

The results <strong>of</strong> the model (table 7.2) were obtained under maximum continuous rating<br />

–6<br />

(MCR). The cost <strong>of</strong> fuel cf was 2.23×<br />

10 $/kJ, <strong>and</strong> the cost <strong>of</strong> water <strong>and</strong> electricity<br />

was also expressed in the most commercial units.<br />

The values are very similar to the results <strong>of</strong> the thermoeconomic model explained<br />

below. This simple model can easily calculate the cost <strong>of</strong> the two main products using<br />

thermoeconomic principles. The only requirement is to introduce the quality <strong>of</strong> the<br />

steam derived to the MSF unit (from the simulator or an owner’s data sheet).<br />

7.1.2 Physical structure<br />

Fuel Product Ch. Equation Cost equation<br />

B<br />

1<br />

C1<br />

= k1<br />

B1<br />

A Branching k *<br />

A<br />

2 Turbine B1<br />

– B2<br />

3 MSF B<br />

2<br />

W<br />

B1<br />

– B2<br />

= k2<br />

D B2<br />

= k3<br />

Results <strong>of</strong> the simple co-generation system model, MCR case.<br />

k*<br />

1 = k 1 cf<br />

The physical structure <strong>of</strong> a plant is similar to a set <strong>of</strong> subsystems or units linked<br />

among themselves <strong>and</strong> to the environment by another set <strong>of</strong> matter, heat, <strong>and</strong> work<br />

that express plant behavior more or less accurately, or, in general:<br />

= k*<br />

1<br />

W k*<br />

2 = k2<br />

D k*<br />

3 = k3<br />

Fuel or product (kW) Unit consumption Exergy cost<br />

C1<br />

= 455,000 k1<br />

= 2.244 k*<br />

1 = 2.244<br />

B1<br />

= 202,800 k2<br />

= 1.293<br />

B2<br />

= 45,000 k3<br />

= 6.553<br />

W = 122,000<br />

D = 6,867<br />

k*<br />

2 = 2.902 (= 0.0233 ($/kWh)<br />

energy system = subsystems or units + matter <strong>and</strong>/or energy flows<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

k *<br />

A<br />

k *<br />

A<br />

k*<br />

3<br />

3 =14.7 (=151.4 kJ/kg, 0.3377 $/m )<br />

(7.1)


FIGURE 7.2<br />

<strong>Thermoeconomic</strong> model<br />

The relationship between the flows <strong>and</strong> subsystems can be set up in a matrix<br />

formulation (Lozano <strong>and</strong> Valero, 1993; Valero et al., 1993), that describes the<br />

balances <strong>of</strong> matter, energy <strong>and</strong> exergy in a very compact form.<br />

The more detailed the definition <strong>of</strong> the physical structure, the better the possibilities<br />

<strong>of</strong> analyzing the installation. However, a more detailed physical structure implies<br />

increasing both the number <strong>of</strong> measurements to be taken in a performance test<br />

(temperatures, pressures, mass flow rates…) <strong>and</strong> the complexity <strong>of</strong> calculations. The<br />

goal is to find an optimum level <strong>of</strong> aggregation, i.e. level <strong>of</strong> detailed description in the<br />

physical structure corresponding to the depth <strong>of</strong> <strong>analysis</strong>.<br />

The physical structure <strong>of</strong> the power plant analyzed thermoeconomically is very<br />

similar to the mathematical model in the simulator (chapter 5). The thermophysical<br />

properties <strong>of</strong> the main flowstreams calculated in a <strong>simulation</strong> can be used as<br />

reasonable initial values for a thermoeconomic <strong>analysis</strong> in an operating condition.<br />

Only the gl<strong>and</strong> steam leakage flow is neglected, which is not significant. Figure 7.2<br />

shows the physical structure <strong>of</strong> the power plant. If the power plant is working in<br />

parallel or twin-extraction mode (that is, the reducing pressure extraction is working),<br />

the pressure reduction station is included in the physical model. Table 7.3 describes<br />

the nomenclature adopted in the previous figure.<br />

Physical structure <strong>of</strong> the power plant considered for the thermoeconomic model.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

163


TABLE 7.3<br />

164<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

Description <strong>of</strong> components appearing in figure 7.2.<br />

Component no. Initials Description<br />

1 CP Condenser Pump<br />

2 LPH2 Low Pressure Heater No. 2<br />

3 LPH1 Low Pressure Heater No. 1<br />

4 DRT Deaerator<br />

5 FP Feed Pump<br />

6 HPH2 High Pressure Heater No. 2<br />

7 HPH1 High Pressure Heater No. 1<br />

8 VEX4 th 4 extraction valve<br />

9 VEX3 rd 3 extraction valve<br />

10 VEXD Extraction valve to deaerator<br />

11 VEX2 nd 2 extraction valve<br />

12 VEX1 st 1 extraction valve<br />

13 VF Feed valve<br />

14 BOI Boiler<br />

15 VB Boiler valve<br />

16 VST Stop valve<br />

17 BHP Brine heater pump<br />

18 HPT1 st<br />

High pressure turbine (1 section)<br />

19 HPT2 nd<br />

High pressure turbine (2 section)<br />

20 HPT3 rd<br />

High pressure turbine (3 section)<br />

21 HPT4 th<br />

High pressure turbine (4 section)<br />

22 LPT1 st<br />

High pressure turbine (1 section)<br />

23 LPT2 nd<br />

High pressure turbine (2 section)<br />

24 CND Condenser<br />

25 GEN Generator<br />

26 MSF Desalination unit (multistage flash)<br />

27 VS1 st<br />

1 Reducing pressure valve (steam)<br />

28 VS2 nd 2 Reducing pressure valve (vac.)<br />

29 VS3 rd<br />

3 Reducing pressure valve (FP)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 7.3<br />

TABLE 7.4<br />

<strong>Thermoeconomic</strong> model<br />

However, the physical model considered for the thermoeconomic <strong>analysis</strong> <strong>of</strong> the<br />

MSF unit (figure 7.3) differs from the mathematical model implemented in the<br />

simulator. The physical model treats the recovery <strong>and</strong> reject sections as a unique<br />

component. If these sections are divided into stages, a huge productive structure is<br />

generated in the plant. Since the functionality <strong>of</strong> each stage is identical, this<br />

possibility <strong>of</strong> plant dissagregation is not considered. Consequently, the input/output<br />

values <strong>of</strong> the recovery <strong>and</strong> reject sections in the simulator can be used as the basis <strong>of</strong><br />

the <strong>analysis</strong> (their operating data are also available). The pump system <strong>of</strong> the<br />

distillation unit is also considered. Exit conditions <strong>of</strong> these pumps are calculated in<br />

the thermoeconomic model with their characteristic charts.<br />

Physical structure <strong>of</strong> the MSF plant considered for the thermoeconomic <strong>analysis</strong>.<br />

Table 7.4 further describes the meaning <strong>of</strong> figure 7.3.<br />

Components description from figure 7.3. Note that component no. 1 is not described in physical<br />

model but is included in other schemes.<br />

Component no. Initials Description<br />

2 BH Brine heater<br />

3 RP Recycle brine pump<br />

4 BDP Blowdown pump<br />

5 RCS Recovery section<br />

6 MIX Mixer<br />

7 RJS Reject section<br />

8 SWP Seawater pump<br />

9 DP Distillate pump<br />

10 MXT Mixer (temper water)<br />

11 TP Tempering pump<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

165


166<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

7.1.3 Productive structure<br />

A plant is more than a set <strong>of</strong> flows <strong>and</strong> units. Each unit has a particular productive<br />

function which contributes to final production. We will clearly indicate which flow or<br />

combination <strong>of</strong> flows constitute the product <strong>of</strong> the unit (P), which ones are the<br />

resources or fuel consumed (F) <strong>and</strong> which flows are the losses (L), i.e. those that<br />

leave the unit <strong>and</strong> plant <strong>and</strong> are not subsequently used.<br />

The productive structure contains the mathematical definition <strong>of</strong> the function <strong>of</strong> each<br />

component. The production objective (product) <strong>and</strong> the resources needed (fuel) to<br />

develop its function are defined for each device, which is equivalent to defining<br />

efficiency. The productive structure also includes the distribution <strong>of</strong> consumed<br />

resources in the different units <strong>and</strong> how plant products are obtained.<br />

The best F-P-L definition to represent unit productive function is obtained by<br />

simultaneously examining their own energy transformation. Using the F-P-L<br />

definition <strong>and</strong> the data from the design <strong>and</strong> operation, it is possible to carry out the<br />

energy <strong>and</strong> exergy <strong>analysis</strong> <strong>of</strong> the plant.<br />

The productive structure can be explained in a diagram with squares representing<br />

physical plant units (productive <strong>and</strong> dissipative physical processes), <strong>and</strong> circles <strong>and</strong><br />

rhombuses that are not physical components <strong>of</strong> the plant. The lines connecting the<br />

different productive units are exergy resources (fuels <strong>and</strong> products). The inlet arrows<br />

going into squares are the fuels <strong>of</strong> the corresponding components <strong>and</strong> outlet arrows<br />

represent products. The circles are branching points where the exergy resource is<br />

distributed to other components. In every junction (rhombus), a significant exergy<br />

resource is obtained by the addition <strong>of</strong> others <strong>of</strong> the same nature but different origin.<br />

To apply an on-line thermoeconomic <strong>analysis</strong> (as presented in Chapter 7) to the dual<br />

plant, the thermoeconomic model should be disaggregated to a deep enough decision<br />

level to make use <strong>of</strong> the most important data provided by the data acquisition system.<br />

The data acquisition system <strong>of</strong> the plant is clearly insufficient to provide the data<br />

required by the productive structure defined in section 7.1 for the power <strong>and</strong><br />

desalination plant. For this reason all required data were provided by the model<br />

presented in chapters 3 to 5, as if they were measured data provided by the plant<br />

acquisition system.<br />

7.1.3.1 Steam power plant<br />

Depending on the <strong>analysis</strong>, a productive structure can be designed in different detail<br />

or aggregation levels.<br />

For instance, in a thermoeconomic <strong>analysis</strong> <strong>of</strong> a power plant,<br />

the MSF plant is considered a single plant unit in the productive structure.<br />

The minimum aggregation level is considered for the MSF plant in the productive<br />

structure <strong>of</strong> the power plant. The F-P definition used for the power plant follows the<br />

trend adopted in conventional steam power plants. The difference between thermal,<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 7.4<br />

<strong>Thermoeconomic</strong> model<br />

mechanical <strong>and</strong> chemical exergy was not considered in the power <strong>and</strong> desalination<br />

plant when the productive structure <strong>of</strong> the system was built. However, a lower<br />

aggregation level was used in the thermoeconomic <strong>analysis</strong> <strong>of</strong> the MSF unit with<br />

several plant units. The F-P-L definition <strong>of</strong> the steam power plant components is<br />

presented in figure 7.4, where B is the exergy flow <strong>of</strong> a stream (its mass flow rate m<br />

multiplied by its specific exergy b), W is the work consumed or generated in a<br />

component, DB is the exergy flow <strong>of</strong> fresh water leaving the MSF, S is the entropy<br />

flow <strong>of</strong> a stream (mass flow m multiplied by the specific entropy s). Exergy losses (L)<br />

are considered but do not explicitly appear in the productive structure.<br />

F-P description in steam power plant.<br />

The fuel <strong>and</strong> product <strong>of</strong> each device is defined depending on the functionality <strong>of</strong> the<br />

component (Frangopoulos, 1990). Thus, the heater is a component installed to heat<br />

feedwater (B4<br />

– B1)<br />

in a Rankine cycle, with extracted steam supplied by the turbine,<br />

which is condensed inside the heater (B2<br />

– B5).<br />

If the heater has a drain from another<br />

heater, the fuel also incorporates its exergy flow (B3).<br />

The job <strong>of</strong> a steam turbine is to<br />

produce work (W) by exhausting the steam from a boiler (B1<br />

– B2).<br />

A pump has the<br />

inverse functionality: it uses work (W) as the fuel to increase the pressure <strong>of</strong> a fluid<br />

(B2<br />

– B1).<br />

A generator is an energy converter, therefore, the fuel is the primary<br />

(mechanical) energy (W1)<br />

<strong>and</strong> the product is secondary (electrical) (W2).<br />

A valve is a<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

167


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

dissipative component. It undergoes exergy losses when the fuel (B 2) passes through<br />

the valve (B 1). The fuel <strong>and</strong> product in a boiler are very clear. A boiler uses primary<br />

energy like natural gas (B gas) to boil <strong>and</strong> superheat the feedwater in a steam cycle<br />

(B 2 -- B 1). The deaerator is a heater with a mixing process <strong>of</strong> several flows. The<br />

product is the heating <strong>of</strong> the colder streams ((m 1 + m 2) b 3 – m 1 b 1 – m 2 b 2) <strong>and</strong> the<br />

fuel is the heat released by hotter streams (m 4 b 4 + m 5 b 5 – (m 4 + m 5) b 3).<br />

The condenser is a dissipative unit which condensates the steam coming from the<br />

steam turbines to produce liquid water. The heat released (Q) has a low temperature<br />

<strong>and</strong> is thus rejected to the atmosphere without any further application. From a<br />

thermodynamic point <strong>of</strong> view, the condenser function allows the working fluid<br />

(water) to reach the physical conditions to perform a new thermodynamic cycle. For<br />

this reason, several authors (Frangopoulos, 1983; Von Spakovsky, 1986; Benelmir,<br />

1989) propose negentropy as the condenser product. The negentropy is a<br />

thermodynamic function (Frangopoulos, 1983) with exergy or energy dimensions but<br />

with entropy reduction <strong>of</strong> water/steam in the condenser. The water/steam entropy is<br />

increased in other plant components. As a result, their negentropy consumption is<br />

primarily produced in the condenser. The amount <strong>of</strong> negentropy consumed in a<br />

component is proportional to its entropy increase. In summary, the exergy losses <strong>of</strong><br />

the different flows entering the condenser are the fuel <strong>of</strong> the device (B 1 + B 2 + B 3 –<br />

B 4). The negentropy produced is the condenser product (T 0 (S 4 – S 3 – S 2 – S 1)).<br />

Finally, the MSF is treated as a component whose main purpose is to produce<br />

freshwater (DB) using different flows <strong>of</strong> steam <strong>and</strong> electricity (B 1 + B 2 – B 3 + W). As<br />

the steam (B 3) is condensed in the heater <strong>of</strong> the distillation unit, some negentropy is<br />

generated in this process (S) which is a secondary product <strong>of</strong> the MSF (auxiliary<br />

product or byproduct).<br />

From the point <strong>of</strong> view <strong>of</strong> the diagnosis, the selected productive structure is<br />

independent <strong>of</strong> the final results (Valero et al., 1999). The maximum aggregation level<br />

provides the product <strong>and</strong> fuel formation cost <strong>of</strong> each component in the steam plant.<br />

Although it is complicated to construct a productive structure with a maximum<br />

aggregation level, it provides the best information to underst<strong>and</strong> the behavior <strong>of</strong> the<br />

individual components <strong>of</strong> a power plant.<br />

The productive structure is made up <strong>of</strong> components with exergy added to the working<br />

fluid <strong>of</strong> the power plant (steam/water). In this case, the components <strong>of</strong> the exergy<br />

addition are the boiler, heaters, deaerator <strong>and</strong> pumps. The amount <strong>of</strong> exergy supplied<br />

to the working fluid is added in a junction <strong>and</strong> then redistributed (using branching<br />

points) to the components where the exergy is removed from the working fluid to be<br />

mixed with another flow or used as fuel <strong>of</strong> a component. The components <strong>of</strong> exergy<br />

removal in the steam cycle with co-generation are the turbine sections, the condenser,<br />

the MSF unit <strong>and</strong> the pressure losses in tubes. Finally, a junction is settled to pick up<br />

the work produced in the turbine sections <strong>and</strong>, after passing the generator, is<br />

redistributed to the components that need the electrical consumption as fuel (pump or<br />

168 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> model<br />

MSF unit). Only two streams leave the plant: distillate flow (DB) <strong>and</strong> net output<br />

power, <strong>and</strong> one stream enters (the exergy flow <strong>of</strong> fuel).<br />

Different productive structures were defined for each operating mode because<br />

different plant units depend on it. For instance, the F-P formulation applied to the<br />

productive structure generated for the more realistic mode, generating power <strong>and</strong><br />

fresh water (extraction mode, see figure 7.5) does not use the live steam reducing<br />

pressure station.<br />

FIGURE 7.5 Productive structure <strong>of</strong> the power plant in extraction mode.<br />

When the power plant is working in condensing mode (only electricity is produced),<br />

brine heater pump <strong>and</strong> MSF components must be removed, <strong>and</strong> consequently, the J3<br />

junction. Figure 7.6 shows the small changes needed to perform the productive<br />

structure <strong>of</strong> the condensing mode.<br />

When the power plant is working in extraction mode at low loads, the low-pressure<br />

turbine is acting as a compressor. As a result, the condenser <strong>and</strong> 2 nd section <strong>of</strong> the<br />

low-pressure turbine are treated as a component with two fuels: work needed to move<br />

the turbine <strong>and</strong> the exergy flow lost in the condenser. Figure 7.7 shows the changes<br />

applied to this operating mode with respect to the first structure (figure 7.5).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 169


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

FIGURE 7.6 Changes applied to extraction mode productive structure (figure 7.5) when the plant operates in<br />

condensing mode.<br />

BHP<br />

17<br />

J3<br />

Vacuum<br />

Steam to MSF<br />

When the power production is less than a minimum (the outlet pressure <strong>of</strong> the fourth<br />

section <strong>of</strong> the high-pressure turbine is very low), the reduction pressure station is<br />

automatically opened to maintain steam conditions to the MSF heater (this is the<br />

parallel mode). The productive structure in figure 7.7 includes the reduction pressure<br />

valve. Figure 7.8 shows the additional structure added to figure 7.5, which is also<br />

valid for the twin extraction mode.<br />

FIGURE 7.7 Productive structure corresponding to extraction mode with low energy production in a dualpurpose<br />

plant. Changes with respect to figure 7.5.<br />

Finally, in desalination or twin desalination mode (steam power plant not working),<br />

the productive structure is quite simple because only six components need to be<br />

considered to perform the productive structure (see figure 7.9), i.e., those operating in<br />

this mode.<br />

170 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

MSF<br />

26


<strong>Thermoeconomic</strong> model<br />

FIGURE 7.8 Productive structure <strong>of</strong> the steam power plant in parallel <strong>and</strong> twin extraction mode. Changes with<br />

respect to figure 7.5.<br />

FIGURE 7.9 Productive structure <strong>of</strong> the steam power plant in desalination or twin desalination mode.<br />

7.1.3.2 MSF unit<br />

The F-P-L definition <strong>of</strong> the MSF components is the first step in building the<br />

productive structure, depending on the aggregation level used to solve the<br />

thermoeconomic model. In this case, recovery <strong>and</strong> reject sections are considered one<br />

component, independently <strong>of</strong> the number <strong>of</strong> their stages. This case could be<br />

considered an intermediate aggregation level, following the physical model<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 171


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

previously defined in figure 7.3. Figure 7.10 resumes the F-P-L definition applied to<br />

the MSF plant units. For more information <strong>of</strong> brine exergy calculation see Annex 2.<br />

The recovery <strong>and</strong> reject sections are complex devices. Their products are very clear:<br />

the distillate produced (DB or DB 2 – DB 1) in each distiller. The resources consumed<br />

are the exergy released by the flashing brine, which is partially recovered by the<br />

cooling brine ((B 1 – B 2) – (F 2 – F 1)), <strong>and</strong> the steam consumed to hold the distillers<br />

below atmospheric pressure (vacuum). Distillate from the recovery section (DB 1) is<br />

also a fuel component <strong>of</strong> the reject section. The brine heater gives the final heating to<br />

the brine (B 4 – B 3) by condensing vapor bled from the turbine (B 1 – B 2). The mixer<br />

device produces an outlet stream (B 3) by merging two or more inlet streams<br />

(B 1 + B 2).<br />

FIGURE 7.10 F-P definition in the MSF unit.<br />

172 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> model<br />

Some other interpretations <strong>of</strong> the F-P definitions were considered to select the<br />

appropriate productive structure. The objective was to obtain the exact value <strong>of</strong> the<br />

exergy cost <strong>of</strong> the final product (whose value is independent <strong>of</strong> the productive<br />

structure) <strong>and</strong> the F-P definition. But the exergy cost <strong>of</strong> the intermediate flowstreams<br />

is obviously different when the fuel <strong>and</strong> product definition <strong>of</strong> each component <strong>and</strong>/or<br />

the aggregation level is changed. The most important point is to find out the physical<br />

sense <strong>of</strong> the flowstreams in the productive structure, in order to explain <strong>and</strong> study the<br />

exergy cost <strong>of</strong> each flow. Several productive structures were studied in this thesis.<br />

• One possibility is to consider that the exergy recovered in the cooling brine<br />

(F 2 -- F 1) is a component <strong>of</strong> the product <strong>of</strong> these components. The fuel <strong>of</strong> these<br />

sections is the exergy released by the flashing brine (B 1 – B 2) <strong>and</strong> the product is<br />

the two effects obtained in the sections (F 2 – F 1) + DB. This results <strong>of</strong> this<br />

definition are similar to the final F-P definition chosen but it contradicts the<br />

functionality <strong>of</strong> the components.<br />

• The heated cooling brine could be considered a subproduct <strong>of</strong> the recovery <strong>and</strong><br />

reject sections while maintaining the fuel as in the previous case. The high value<br />

<strong>of</strong> the subproduct (F 2 – F 1), (several times the value <strong>of</strong> distilled water in these<br />

sections) gives nonsense values for the calculated exergy costs.<br />

• Consider a zero exergy cost <strong>of</strong> the MSF plant residues (fourth proposition <strong>of</strong> the<br />

exergy cost theory, Valero et al., 1986a). The cost <strong>of</strong> the rejected cooling<br />

seawater <strong>and</strong> blowdown is not charged over the rest <strong>of</strong> the MSF plant<br />

flowstreams. This avoids introducing the fictitious device in the productive<br />

structure <strong>of</strong> the distillation plant. This consideration is a price allocation because<br />

the residues are final products external to the system <strong>and</strong> have zero cost.<br />

• The distilled water in the recovery section may not be considered a fuel <strong>of</strong> the<br />

reject section. The product <strong>of</strong> the reject section should only be the quantity <strong>of</strong><br />

distilled water produced in that section, not the total amount <strong>of</strong> freshwater<br />

produced. This scenario only varies the cost <strong>of</strong> reject section.<br />

• The system recovery-reject section could be considered a component, in order to<br />

avoid the effect <strong>of</strong> recycling flows in the MSF plant <strong>and</strong> the modeling <strong>of</strong> a<br />

fictitious mixer in the final stage <strong>of</strong> the distillation plant. This is a higher<br />

aggregation level than adopted in this thesis.<br />

• It would not be adequate to consider the chemical exergy <strong>of</strong> the distillate leaving<br />

the reject section as the final product <strong>of</strong> the MSF plant. Its low value would<br />

imply huge exergy operating costs <strong>of</strong> the rest <strong>of</strong> the flowstreams inside the<br />

distillers. Furthermore, the <strong>analysis</strong> <strong>of</strong> a thermal inefficiency in a distiller cannot<br />

be performed with the F-P definition adopted in this hypothetical assumption.<br />

The chemical exergy <strong>of</strong> freshwater only depends on salt concentration <strong>and</strong> does<br />

not vary under thermal inefficiency. The only consequence <strong>of</strong> a thermal<br />

inefficiency is a thermal exergy variation.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 173


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

The formation <strong>of</strong> the productive structure <strong>of</strong> the MSF unit is not easily explained with<br />

the F-P definition considered in the thermoeconomic model (several junctions are<br />

needed to obtain component fuel <strong>and</strong> product). As the flows circulating by the MSF<br />

unit are pumped, the main flows <strong>of</strong> the plant are added to a junction in which the<br />

exergy added by the pump is incorporated to the flow. The most significant branching<br />

points <strong>of</strong> the MSF plant redistribute their product as a fuel for some components <strong>of</strong><br />

the MSF unit. The first one is the cooling brine heated in the brine heater, the second<br />

branching has the cooling seawater to reject. But the most amazing situation <strong>of</strong> this<br />

structure is the non-physical component or fictitious device (FD). It was included at<br />

the beginning <strong>of</strong> the structure to account for residue costs (blowdown <strong>and</strong> reject<br />

cooling seawater) in the thermoeconomic model. The cost <strong>of</strong> steam to brine heater<br />

(considered to be the main fuel <strong>of</strong> the plant) is overcharged by the effect <strong>of</strong> the two<br />

useless flows sent to sea. The exergy costs <strong>of</strong> the blowdown <strong>and</strong> discharged cooling<br />

brine are used <strong>and</strong> conveniently incorporated into the rest <strong>of</strong> the internal costs <strong>and</strong> the<br />

final product <strong>of</strong> the MSF unit.<br />

Figure 7.11 shows the productive structure <strong>of</strong> the MSF plant corresponding to the F-P<br />

definition explained above (figure 7.10), the number <strong>of</strong> junctions <strong>and</strong> branches are a<br />

result <strong>of</strong> the F-P definition adopted for the recovery <strong>and</strong> reject sections. The operating<br />

modes <strong>of</strong> the power plant do not affect the productive structure <strong>of</strong> the desalination<br />

unit, unless the condensing mode is selected (in this case there is no freshwater<br />

production).<br />

FIGURE 7.11 Productive structure <strong>of</strong> the MSF unit.<br />

174 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> model<br />

7.1.4 <strong>Thermoeconomic</strong> model<br />

The thermoeconomic model is the mathematical representation <strong>of</strong> the productive<br />

structure. It consists <strong>of</strong> a group "characteristic equations" which express (for all<br />

components in the productive structure) each inlet flow as a function <strong>of</strong> the outlet<br />

flows <strong>and</strong> a set <strong>of</strong> internal parameters, i.e.:<br />

Unit j: F i = κ ij · P j (7.2)<br />

Junction j: F i = r ij · P j (7.3)<br />

Branching point j: F j = ∑ P i (7.4)<br />

where κ is the technical production coefficient <strong>of</strong> the unit <strong>and</strong> r is a structural<br />

parameter in the junctions or exergy ratio. Equation (7.2) provides information about:<br />

• the productive function <strong>of</strong> each commoponent, i.e. its production (P),<br />

• what the component needs (F) to develop its productive purpose, <strong>and</strong><br />

• the thermodynamic efficiency (κ) <strong>of</strong> the process taking place in the component.<br />

The structural equations (7.3) <strong>and</strong> (7.4) contain the distribution <strong>of</strong> the resources<br />

consumed by the plant components, i.e. how the components are interconnected from<br />

a productive viewpoint.<br />

The <strong>Thermoeconomic</strong> model <strong>of</strong> the steam power plant (extraction mode) has one 7.2type<br />

equation for each fuel entering a component (57 equations in total), four<br />

equations for the four junctions <strong>and</strong> four equations derived from the four branching<br />

points in the productive structure (figure 7.5). There are 19 characteristic equations in<br />

the MSF unit model <strong>and</strong> seven <strong>and</strong> three equations corresponding to the junctions <strong>and</strong><br />

branching points.<br />

The characteristic equations (equations 7.2–7.4) can easily be written using the<br />

productive structure diagram. The subscript numbers <strong>of</strong> the fuel <strong>and</strong> products<br />

correspond to the flow diagram <strong>of</strong> Chapter 4 (power plant scheme) <strong>and</strong> Chapter 3<br />

(diagram <strong>of</strong> the MSF plant). Table 7.5 includes the equations that describe the<br />

thermoeconomic model <strong>of</strong> the steam power plant.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 175


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

TABLE 7.5 Exergy flows <strong>and</strong> characteristic equations <strong>of</strong> components in the steam power plant (extraction<br />

mode).<br />

Dev. Exergy Flows Characteristic equation(s)<br />

CP<br />

PCP FSCP = m12 (b12 –b11 )<br />

= m12 T0 (s12 –s11 )<br />

WCP FSCP = kBCP * PCP = kSCP * PCP LPH2<br />

LPH1<br />

DRT<br />

FP<br />

HPH2<br />

HPH1<br />

VEX4<br />

VEX3<br />

VEXD<br />

PLPH2 = m12 (b14 –b12 )<br />

FB1LPH2 = m34 (b34 –b25 )+ mci (bci –b25 )<br />

FB2LPH2 = m33 (b23 –b25 )<br />

FS LPH2<br />

= T 0 {m 12 (s 14 –s 12 ) – m 34 (s 34 –s 25 )<br />

– m 33 (s 33 –s 25 )– m ci (s ci –s 25 )}<br />

PLPH1 = m12 (b15 –b14 )<br />

FBLPH1 = m33 (b33 –b23 )<br />

FSLPH1 PDRT = T 0 {m 12 (s 15 –s 14 )– m 33 (s 33 –s 23 )}<br />

= m12 (b16 –b15 )+ mdes (b16 –brdes )<br />

FB1DRT = m32 (b32 –b16 )<br />

FB2DRT = (m30 + m31 ) (b22 –b16 )<br />

FSDRT = T0 {m20 s16 –(m30 + m31 ) s22 –m12 s15 – m32 s32 – mdes srdes} PFP FSCP = m20 (b17 –b16 )<br />

= m20 T0 (s17 –s16 )<br />

PHPH2 = m20 (b19 –b17 )<br />

FB1HPH2 = m31 (b31 –b22 )<br />

FB2HPH2 = m30 (b21 –b22 )<br />

FS HPH2 = T 0 {m 20 (s 19 –s 17 )–m 31 (s 31 –s 22 )<br />

– m 30 (s 21 –s 22 )}<br />

PHPH1 = m20 (b20 –b19 )<br />

FBHPH1 = m30 (b30 –b21 )<br />

FSHPH1 = T0 {m20 (s20 –s19 )–m30 (s30 –s21 )}<br />

PVEX4 = m34 (b34 –b25 ) + mci (bci –b25 )<br />

FBVEX4 = m34 (b8 –b25 ) + mci (bci –b25 )<br />

FSVEX4 = T0 m34 (s34 –s8 )<br />

PVEX3 = m33 (b33 –b23 )<br />

FBVEX3 = m33 (b6 –b23 )<br />

FSVEX3 = T0 m33 (s33 –s6 )<br />

PVEXD = m32 (b32 –b16 )<br />

FBVEXD = m32 (b5 –b16 )<br />

FSVEXD = T0 m32 (s32 –s5 )<br />

FB1LPH2 = kB1LPH2 * PLPH2 FB2LPH2 = kB2LPH2 * PLPH2 FSLPH2 = kSLPH2 * PLPH2 FBLPH1 = kBLPH1 * PLPH1 FSLPH1 = kSLPH1 * PLPH1 FB1DRT = kB1DRT * PDRT FB2DRT = kB2DRT * PDRT FSDRT = kSDRT * PDRT 176 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

W FP<br />

FS FP<br />

= kB FP * P FP<br />

= kS FP * P FP<br />

FB1HPH2 = kB1HPH2 *PHPH2 FB2HPH2 = kB2HPH2 * PHPH2 FSHPH2 = kSHPH2 * PHPH2 FBHPH1 = kBHPH1 * PHPH1 FSHPH1 = kSHPH1 * PHPH1 FB VEX4 = kB VEX4 * P VEX4<br />

FS VEX4 = kS VEX4 * P VEX4<br />

FB VEX3 = kB VEX3 * P VEX3<br />

FS VEX3 = kS VEX3 * P VEX3<br />

FB VEXD = kB VEXD * P VEXD<br />

FS VEXD = kS VEXD * P VEXD


VEX2<br />

VEX1<br />

VF<br />

BOI<br />

VB<br />

<strong>Thermoeconomic</strong> model<br />

TABLE 7.5 Exergy flows <strong>and</strong> characteristic equations <strong>of</strong> components in the steam power plant (extraction<br />

mode).<br />

Dev. Exergy Flows Characteristic equation(s)<br />

VST<br />

BHP<br />

HPT1<br />

HPT2<br />

HPT3<br />

HPT4<br />

LPT1<br />

LPT2<br />

CND<br />

PVEX2 = m31 (b31 –b22 )<br />

FBVEX2 = m31 (b4 –b22 )<br />

FSVEX2 = T0 m31 (s31 –s4 )<br />

PVEX1 = m30 (b30 –b21 )<br />

FBVEX1 = m30 (b3 –b21 )<br />

FSVEX1 = T0 m30 (s30 –s3 )<br />

PVF FBVF FSVF PBOI FSBOI PVB FBVB FSVB PVST FBVST FSVB PBHP FSBHP = m12 (b28 –b11 ) + (m20 –m12 ) (b28 –b16 )<br />

= m12 (b20 –b11 ) + (m20 –m12 ) (b20 –b16 )<br />

= T0 m20 (s28 –s20 )<br />

= m20 (b29 –b28 )<br />

= T0 m20 (s29 –s28 )<br />

= m12 (b1 –b11 ) + (m20 –m12 ) (b1 –b16 )<br />

= m12 (b29 –b11 ) + (m20 –m12 ) (b29 –b16 )<br />

= T0 m20 (s1 –s29 )<br />

= m12 (b1’ –b11 ) + (m20 –m12 ) (b1’ –b16 )<br />

= PVB = T0 m20 (s1’ –s1 )<br />

= mdes (brdes –bdes )<br />

= T0 mdes (srdes –sdes )<br />

FBHPT1 = m20 (b1’ –b3 )<br />

= T0 m20 (s3 –s1’ )<br />

FS HPT1<br />

FBHPT2 = (m20 –m30 –mva ) (b3 –b4 )<br />

= T0 (m20 –m30 –mva ) (s4 –s3 )<br />

FS HPT2<br />

FBHPT3 = (m20 –m30 –mva–m31 ) (b4 –b5 )<br />

= T0 (m20 –m30 –mva –m31 ) (s5 –s4 )<br />

FS HPT3<br />

FBHPT4 = (m20 –m30 –mva –m31 –m32 ) (b5 –b6 )<br />

= T0 (m20 –m30 –mva –m31 –m32 ) (s6 –s5 )<br />

FS HPT4<br />

FBLPT1 FSLPT1 FBLPT2 FSLPT2 P CND<br />

FB CND<br />

= (m9 + m34 ) (b6 –b8 )<br />

= T0 (m9 + m34 ) (s8 –s6 )<br />

= m9 (b8 –b9 )<br />

= T0 m9 (s9 –s8 )<br />

= T0 {m9 s9 + (m34 + m33 + mci ) s25 + mva sva –m12 s11 }<br />

= m9 b9 + (m34 + m33 + mci ) b25 + mva bva – m12 b11 FB VEX2 = kB VEX2 * P VEX2<br />

FS VEX2 = kS VEX2 * P VEX2<br />

FB VEX1 = kB VEX1 * P VEX1<br />

FS VEX1 = kS VEX1 * P VEX1<br />

FB VF<br />

FS VF<br />

C 1<br />

FS BOI<br />

FB VB<br />

FS VB<br />

FB VST<br />

FS VST<br />

W BHP<br />

FS BHP<br />

= kB VF * P VF<br />

= kS VF * P VF<br />

= kB BOI * P BOI<br />

= kS BOI * P BOI<br />

= kB VB * P VB<br />

= kS VB * P VB<br />

= kB VST * P VST<br />

= kS VST * P VST<br />

= kB BHP * P BHP<br />

= kS BHP * P BHP<br />

FB HPT1 = kB HPT1 * W HPT1<br />

FS HPT1<br />

= kS HPT1 * W HPT1<br />

FB HPT2 = kB HPT2 * W HPT2<br />

FS HPT2<br />

= kS HPT2 * W HPT2<br />

FB HPT3 = kB HPT3 * W HPT3<br />

FS HPT3<br />

= kS HPT3 * W HPT3<br />

FB HPT4 = kB HPT4 * W HPT4<br />

FS HPT4<br />

FB LPT1<br />

FS LPT1<br />

FB LPT2<br />

FS LPT2<br />

FB CND<br />

= kS HPT4 * W HPT4<br />

= kB LPT1 * W LPT1<br />

= kS LPT1 * W LPT1<br />

= kB LPT2 * W LPT2<br />

= kS LPT2 * W LPT2<br />

= kB CND * P CND<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 177


GEN<br />

MSF<br />

W T<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

TABLE 7.5 Exergy flows <strong>and</strong> characteristic equations <strong>of</strong> components in the steam power plant (extraction<br />

mode).<br />

= W HPT1 + W HPT2 + W HPT3<br />

+ W HPT4 + W LPT1 + W LPT2<br />

FB1MSF = mdes (b6 –bdes )<br />

FB2MSF = mva (b3 –bva )<br />

FS MSF<br />

A FB J3 = m des (b 6 –b 16 )<br />

= T 0 {m des (s des –s 6 )+ m va (s va –s 3 )}<br />

The physical model <strong>of</strong> the thermoeconomic <strong>analysis</strong> differs from the mathematical<br />

model presented in Chapter 3. Figure 7.12 shows the exergy flows considered in the<br />

thermoeconomic model <strong>of</strong> the MSF plant (which also appear in the characteristic<br />

equations in table 7.6). We used the flow nomenclature adopted in Chapter 3.<br />

FIGURE 7.12 Physical model considered in the thermoeconomic <strong>analysis</strong> <strong>of</strong> the MSF plant.<br />

178 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

W T<br />

= k GEN * P GEN<br />

WMSF = kB3MSF * PD FB1MSF = kB1MSF * PD FB2MSF = kB2MSF * PD FSMSF = kSMSF * PD P VST = FB VEX4 + FB VEX3 + FB VEX2 + FB VEX1<br />

+ FB VEXD + FB2 LPH2 + FB2 DRT + FB2 HPH2<br />

+ FB HPT1 + FB HPT2 + FB HPT3 + FB HPT4<br />

+ FB LPT1 + FB LPT2 + FB J3 + FB CND +FB2 MSF<br />

B P DRT = m 12 (b 16 –b 15 ) + m des (b 16 –b rdes )<br />

C P GEN = W TN + W FP + W CP +W MSF +W BHP<br />

J1<br />

Dev. Exergy Flows Characteristic equation(s)<br />

F VF<br />

= r FP * P FP + r LPH2 * P LPH2 + r LPH1<br />

* P LPH1 + r DRTj1 * m 12 (b 16 –b 15 )<br />

+ r CP * P CP + r HPH2 * P HPH2 + r HPH1 * P HPH1<br />

J2 FVB = rVF * PVF + rBOI * PBOI J3<br />

FB1MSF = rJ3 * FBJ3 + rDRTj3 * mdes (b16 –brdes ) + rBHP * PBHP WT = rHPT1 * WHPT1 + rHPT2 * WHPT2 J4<br />

+ rHPT3 * WHPT3 + rHPT4 * WHPT4 + rLPT1 * WLPT1 + rLPT2 * WLPT2


<strong>Thermoeconomic</strong> model<br />

TABLE 7.6 Exergy flows <strong>and</strong> characteristic equations for the components <strong>of</strong> the MSF plant.<br />

Devices Exergy flows Characteristic equation(s)<br />

FD<br />

BH<br />

P FD = m des (b 6 – b des ) ≡ FB1 MSF<br />

F1 FD = P FD<br />

F2 FD = BD b 10<br />

F3 FD = CW b 13<br />

PBH = R (b4 – b3 )<br />

FBH = PFD F1 FD = k1 FD * P FD<br />

F2 FD = k2 FD * P FD<br />

F3 FD = k3 FD * P FD<br />

P BH = k BH * P BH<br />

RP P RP = R (b 7 – b 8 ) W RP = k RP * P RP<br />

BDP P BDP = BD (b 10 – b 8 ) W BDP = k BDP * P BDP<br />

RCS<br />

MIX<br />

RJS<br />

PRCS = Drcs b5 F1RCS = R b4 – (R – Drcs ) b6 – R (b3 – b7 )<br />

F2RCS ≡ 0.5 FB2MSF F3 RCS = 0.5 m vent b 15<br />

P MIX = R b 8<br />

F1 MIX = (R – D – BD) b 9<br />

F2 MIX = F b 13<br />

P RJS = D b 11<br />

F1RJS = (R – Drcs ) b6 – (R – D) b9 + Drcs b5 – SR (b13 – b17 )<br />

F2RJS = 0.5 FB2MSF F3RJS = 0.5 mvent b15 F1 RCS = k1 RCS * P RCS<br />

F2 RCS = k2 RCS * P RCS<br />

F3 RCS = k3 RCS * P RCS<br />

F1MIX = k1MIX * PMIX F2MIX = k2MIX * PMIX F1RJS = k1RJS * PRJS F2RJS = k2RJS * PRJS F3RJS = k3RJS * PRJS SWP P SWP = SW (b 15 – b 16 ) W SWP = k SWP * P SWP<br />

DP P DP = D (b 12 – b 11 ) W DP = k DP * P DP<br />

MXT<br />

P MXT = SR b 17<br />

F1 MXT = TP b 14<br />

F2 MXT = (SW – m vent ) b 15<br />

F1 MXT = k1 MXT * P MXT<br />

F2 MXT = k2 MXT * P MXT<br />

TP P TP = TP (b 14 – b 13 ) W TP = k TP * P TP<br />

JA<br />

P JA = F2 FD<br />

F1 JA = P BDP<br />

F2 JA = BD b 8<br />

P JA = r1 JA * F1 JA + r2 JA * F2 JA<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 179


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

TABLE 7.6 Exergy flows <strong>and</strong> characteristic equations for the components <strong>of</strong> the MSF plant.<br />

Devices Exergy flows Characteristic equation(s)<br />

JB<br />

C<br />

JD<br />

7.2 Cost <strong>analysis</strong><br />

PJB = R b4 – R (b3 – b7 )<br />

F1JB = PBH F2JB = PRP F3JB = PMIX F1JD = (R – Drcs ) b6 – (R – D) b8 – SR (b13 – b17 )<br />

F1JI = SR (b13 – b17 )<br />

P JD = F1 RJS<br />

F2 JD = P RCS<br />

P JB = r1 JB * F1 JB + r2 JB * F2 JB<br />

+ r3 JB * F3 JB<br />

P JB = F2 JA + F1 JD + F1 RCS + F1 JI<br />

+ F1 MIX<br />

P JD = r1 JD * F1 JD + r2 JD * F2 JD<br />

E F1 JK = TP b 13 P JI = F3 FD + F1 JK + F2 MIX<br />

JG<br />

P JG = SR b 15<br />

F1 JG = P SWP<br />

F2 JG = SR b 16<br />

P JG = r1 JG * F1 JG + r2 JG * F2 JG<br />

H P JG = F3 RCS + F3 RJS + F2 MXT<br />

JI<br />

JJ<br />

JK<br />

P JI = SR b 13<br />

F2 JI = P MXT<br />

P JJ = P D = D b 12<br />

F1 JJ = P RJS<br />

F2 JJ = P DP<br />

P JK = F1 MXT<br />

F2 JK = P TP<br />

P JI = r1 JI * F1 JI + r2 JI * F2 JI<br />

P JJ = r1 JJ * F1 JJ + r2 JJ * F2 JJ<br />

P JK = r1 JK * F1 JK + r2 JK * F2 JK<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> combines the First <strong>and</strong> Second Law <strong>of</strong> Thermodynamics<br />

along with monetary cost balances at the system component level. It helps to<br />

underst<strong>and</strong> the process <strong>of</strong> cost formation, minimize overall product costs <strong>and</strong> assess<br />

costs <strong>of</strong> the different products obtained in the processes. The cost accounting method<br />

can calculate costs using rough data from an energy system control room (pressures,<br />

temperatures, mass flow rates, electrical production, fuel consumption, excess <strong>of</strong><br />

oxygen etc. <strong>and</strong> the economic data).<br />

The costs <strong>of</strong> all significant mass <strong>and</strong> energy flowstreams is a very powerful <strong>and</strong><br />

interesting piece <strong>of</strong> information about the amount <strong>of</strong> resources used to obtain each<br />

180 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Cost <strong>analysis</strong><br />

significant mass <strong>and</strong> energy flowstream. Knowing the costs <strong>of</strong> the mass <strong>and</strong> energy<br />

flowstreams is the key to thermoeconomic <strong>analysis</strong>. The first consequence is price<br />

assessment <strong>of</strong> the products based on physical criteria.<br />

7.2.1 Exergy costs allocation<br />

Valero et al. (1986a) present the fundamental problem <strong>of</strong> cost allocation as follows:<br />

Given a system whose limits have been defined <strong>and</strong> a level <strong>of</strong> aggregation that<br />

specifies the subsystems which constitute it, how to obtain the cost <strong>of</strong> all the flows that<br />

become interrelated in this structure.<br />

The origin <strong>of</strong> every cost lies in the irreversibility <strong>of</strong> the processes. This is a<br />

cornerstone in thermoeconomics. But how do we link the variation in the local<br />

irreversibility (∆I i) to the increase <strong>of</strong> resources consumed (∆F T)?<br />

Two factors are added to consider the economic: market prices (cf), which are not<br />

necessarily linked to the exergy <strong>of</strong> the processed resources <strong>and</strong> depreciation, <strong>and</strong><br />

maintenance costs <strong>of</strong> the productive process (Z). The thermoeconomic cost <strong>of</strong> a flow<br />

can be calculated after the second factor is introduced (section 7.2.3). The exergy<br />

costs calculated in this section only take into account the fuel consumed to produce<br />

each flowstream.<br />

Valero et al (1986a) also propose a rational procedure to determine the cost <strong>of</strong> all<br />

mass <strong>and</strong> energy flowstreams based on four propositions presented in the ‘Theory<br />

<strong>of</strong> exergetic cost’. Consider a plant with n units <strong>and</strong> m flows with known exergy<br />

flows. The set <strong>of</strong> balances <strong>of</strong> exergy costs (P1 proposition) <strong>of</strong> the n units provides a<br />

system <strong>of</strong> n equations. The number <strong>of</strong> flows will be higher than the number <strong>of</strong><br />

units, <strong>and</strong> (m – n) auxiliary equations will be needed to determine flow cost. Serra<br />

(1994) demonstrated that the rest <strong>of</strong> the required equations are obtained from the<br />

productive structure <strong>of</strong> the plant through the F-P-L definition <strong>of</strong> its units <strong>and</strong> the<br />

subsequent application <strong>of</strong> the Theory <strong>of</strong> exergetic cost.<br />

The Structural Theory <strong>of</strong> <strong>Thermoeconomic</strong>s (Valero et. al, 1993) based on the rules <strong>of</strong><br />

mathematical derivation provides exactly the same system <strong>of</strong> cost equations.<br />

Consequently, this theory can calculate flow cost <strong>of</strong> the above four propositions by<br />

simply applying the chain rule <strong>of</strong> derivatives to the characteristic equations <strong>of</strong> the<br />

thermoeconomic model (as explained in Chapter 6). The system <strong>of</strong> equations<br />

providing the exergy costs <strong>of</strong> the steam power plant (cost <strong>of</strong> the flows appearing in<br />

the productive structure depicted in figure 7.5) is shown in table 7.7. Note that<br />

negentropy is included in the cost equation <strong>of</strong> each component as a second fuel. The<br />

negentropy generated in the condenser must be allocated to the rest <strong>of</strong> the plant<br />

components as a function <strong>of</strong> their entropy increase.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 181


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

TABLE 7.7 System <strong>of</strong> equations providing the unit exergy costs <strong>of</strong> the steam power plant (extraction mode).<br />

Device Exergy cost balance<br />

*<br />

*<br />

CP kCP = kBCP kCPw + kSCP *<br />

*<br />

*<br />

LPH2 kLPH2 = kB1LPH2 kVEX4 + kB2LPH2 kLPH2v + kSLPH2 *<br />

*<br />

LPH1 kLPH1 = kBLPH1 kVEX3 + kSLPH1 *<br />

*<br />

*<br />

DRT kDRT = kB1DRT kVEXD + kB2DRT kDRTv + kSDRT *<br />

*<br />

FP kFP = kBFP kFPw + kSFP *<br />

*<br />

*<br />

HPH2 kHPH2 = kB1HPH2 kVEX2 + kB2HPH2 kHPH2v + kSHPH2 *<br />

*<br />

HPH1 kHPH1 = kBHPH1 kVEX1 + kSHPH1 *<br />

*<br />

VEX4 kVEX4 = kBVEX4 kVEX4v + kSVEX4 *<br />

*<br />

VEX3 kVEX3 = kBVEX3 kVEX3v + kSVEX3 *<br />

*<br />

VEXD kVEXD = kBVEXD kVEXDv + kSVEXD *<br />

*<br />

VEX2 kVEX2 = kBVEX2 kVEX2v + kSVEX2 *<br />

*<br />

VEX1 kVEX1 = kBVEX1 kVEX1v + kSVEX1 *<br />

*<br />

VF kVF = kBVF kJ1 + kSVF *<br />

*<br />

BOI kBOI = kBBOI kFUEL + kSBOI *<br />

*<br />

VB kVB = kBVB kJ2 + kSVB *<br />

*<br />

VST kVST = kBVST kVB + kSVST *<br />

*<br />

BHP kBHP = kBBHP kBHPw + kSBHP *<br />

*<br />

HPT1 kHPT1 = kBHPT1 kHPT1v + kSHPT1 *<br />

*<br />

HPT2 kHPT2 = kBHPT2 kHPT2v + kSHPT2 *<br />

*<br />

HPT3 kHPT3 = kBHPT3 kHPT3v + kSHPT3 *<br />

*<br />

HPT4 kHPT4 = kBHPT4 kHPT4v + kSHPT4 182 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

*<br />

kCPs *<br />

kFPs *<br />

kVFs *<br />

kVBs *<br />

kBOIs *<br />

kVSTs *<br />

kLPH1s *<br />

kHPH1s *<br />

kBHPs *<br />

kVEX4s *<br />

kVEX3s *<br />

kVEXDs *<br />

kVEX2s *<br />

kVEX1s *<br />

kHPT1s *<br />

kHPT2s *<br />

kHPT3s *<br />

kHPT4s *<br />

kDRTs *<br />

kLPH2s *<br />

kHPH2s


Cost <strong>analysis</strong><br />

TABLE 7.7 System <strong>of</strong> equations providing the unit exergy costs <strong>of</strong> the steam power plant (extraction mode).<br />

Device Exergy cost balance<br />

*<br />

*<br />

LPT1 kLPT1 = kBLPT1 kLPT1v + kSLPT1 *<br />

*<br />

LPT2 kLPT2 = kBLPT2 kLPT2v + kSLPT2 *<br />

CND kCND = kBCND *<br />

GEN kGEN = kBGEN *<br />

*<br />

*<br />

*<br />

MSF kMSF = kB3MSF kMSFw + kB1MSF kJ3 + kB2MSF kMSFv + kSMSF J1<br />

* *<br />

J2 kJ2 = rVF kVF + rBOI *<br />

*<br />

*<br />

J3 kJ3 = rDRTj3 kDRTJ3 + rVA kJ3v + rBHP J4<br />

A<br />

*<br />

*<br />

*<br />

*<br />

= rHPT1 kHPT1 + rHPT2 kHPT2 + rHPT3 kHPT3 + rHPT4 kHPT4 + rLPT1 + r LPT2<br />

* * * *<br />

B kGEN = kFPw = kCPw = kMSFw =<br />

* *<br />

C kDRT = kDRTJ1 =<br />

D<br />

*<br />

kJ1 *<br />

kJ4 *<br />

kCNDv *<br />

kJ4 *<br />

kLPT1s *<br />

kLPT2s *<br />

kMSFs *<br />

*<br />

*<br />

*<br />

*<br />

= rFP kFP + rLPH2 kLPH2 + rLPH1 kLPH1 + rDRTj1 kDRTJ1 + rHPH2 kHPH2 *<br />

*<br />

+ rHPH1 kHPH1 + rCP kCP *<br />

kLPT2 *<br />

kBOI *<br />

kBHP *<br />

kLPT1 * * * *<br />

*<br />

*<br />

*<br />

*<br />

kVST = kLPH2v = kDRTv = kHPH2v = kVEX4v = kVEX3v = kVEX2v = kVEX1v * * * * * * *<br />

= kHPT1v = kHPT2v = kHPT3v = kHPT4v = kLPT1v = kHPT2v = kCNDv * *<br />

= kMSFv = kJ3v *<br />

kDRTJ3 *<br />

kBHPw * * * * * * * *<br />

kCND = kCPs = kLPH1s = kLPH2s = kDRTs = kFPs = kHPH1s = kHPH2s *<br />

*<br />

*<br />

*<br />

*<br />

* *<br />

= kVEX4s = kVEX3s = kVEXDs = kVEX2s = kVEX1s = kVFs = kBOIs * * * * * * *<br />

= kVBs = kVSTs = kBHPs = kHPT1s = kHPT2s = kHPT3s = kHPT4s * * *<br />

= kLPT1s = kLPT2s = kMSFs In the system <strong>of</strong> equations providing the exergy costs <strong>of</strong> the MSF plant (table 7.8), the<br />

negentropy does not appear, although the brine heater is acting as a plant condenser.<br />

The negentropy decreases energy waste in the condenser <strong>and</strong> improves the power<br />

plant efficiency.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 183


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

TABLE 7.8 System <strong>of</strong> equations providing the exergy costs <strong>of</strong> the MSF plant (figure 7.11).<br />

Components Exergy cost equations<br />

*<br />

*<br />

*<br />

FD kFD = k1FD kST + k2FD kJA + k3FD *<br />

BH kBH = kBH *<br />

RP kRP = kRP *<br />

BDP kBDP = kBDP *<br />

*<br />

*<br />

RCS kRCS = k1RCS kRCSf1 + k2RCS kVA + k3RCS *<br />

*<br />

MIX kMIX = k1MIX kMIXf1 + k2MIX *<br />

*<br />

*<br />

RJS kRJS = k1RJS kJD + k2RJS kVA + k3RJS *<br />

SWP kSWP = kSWP *<br />

DP kDP = kDP *<br />

*<br />

MXT kMXT = k1MXT kJK + k2MXT *<br />

TP kTP = kTP *<br />

kFD *<br />

kW *<br />

kW *<br />

kW *<br />

kW *<br />

kW *<br />

*<br />

JA kJA = r1JA kBDP + r2JA *<br />

*<br />

*<br />

JB kJB = r1JB kBH + r2JB kRP + r3JB *<br />

*<br />

JD kJD = r1JD kJDf1 + r2JD *<br />

*<br />

JG kJG = r1JG kSWP + r2JG * *<br />

JI kJI = r1JI kJIf1 + r2JI * *<br />

JJ kJJ = r1JJ kRJS + r2JJ *<br />

*<br />

Jk kJK = r1JK kJKf1 + r2JK * * * * *<br />

C kJB = kJAf2 = kJDf1 = kRCSf1 = kJIf1 =<br />

* * *<br />

E kJI = kMIXf2 = kFDf3 =<br />

*<br />

kJAf2 *<br />

kRCS * * *<br />

F kJG = kRCSf3 = kRJSf3 =<br />

*<br />

kMIXf2 *<br />

kMXTf2 *<br />

kMIX *<br />

kFDf3 *<br />

kRCSf3 *<br />

kRCSf3 184 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

*<br />

kSW *<br />

kMXT *<br />

kDP *<br />

kTP *<br />

kJKf1 *<br />

kMXTf2 *<br />

kMIXf1


Cost <strong>analysis</strong><br />

7.2.2 Exergy cost <strong>analysis</strong><br />

We calculated the exergy costs for the productive structure in figure 7.5 <strong>and</strong> analyzed<br />

them under eight different operating conditions with equations in table 7.7. They are<br />

expressed in energy units <strong>and</strong> represent the amount <strong>of</strong> resources (usually natural gas)<br />

consumed to obtain each significant mass <strong>and</strong> energy flowstream. These only<br />

represent the operation costs (they do not include the cost <strong>of</strong> each plant device) in<br />

terms <strong>of</strong> energy.<br />

The thermodynamic properties <strong>of</strong> the mass <strong>and</strong> energy flowstreams (figures 7.2<br />

<strong>and</strong> 7.3) were obtained by the simulator. The main features <strong>of</strong> each case are shown in<br />

table 7.9. Most <strong>of</strong> them correspond to a performance data case <strong>of</strong> the power plant,<br />

already described in Chapter 4.<br />

TABLE 7.9 Case studies <strong>of</strong> the exergy cost <strong>analysis</strong> (PTC: Performance Test Case <strong>of</strong> the dual plant; Gc:<br />

Natural gas consumed; CBS: Cleaning Ball System was used).<br />

Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 Case 7 Case 8<br />

PTC MR ODOB MCR MSL4 PL85 — MSL3 —<br />

W (kW) 146,693 — 122,000 75,440 91,000 53,500 76,500 71,000<br />

mls (kg/s) 156.187 70.38 156.187 109.5 117.39 70.0 170.0 160.0<br />

Gc (Nm 3 /h) 43,090 22,780 43,460 31,560 33,650 20,850 49,340 49,390<br />

LS (GCal/h) 0.0 150.0 0.0 0.0 0.0 0.0 150.0 150.0<br />

mdes (kg/s) 0.0 88.5 89.68 88.63 75.62 41.7 83.0 73.5<br />

Pc (bar) 0.135 — 0.072 0.021 0.055 0.048 0.048 0.048<br />

D (T/h) — 2,418.0 2,418.0 2,418.0 2,060.0 1,216.3 2,260.5 2,309.5<br />

TBT (º C) — 112.0 112.0 112.0 100.0 84.0 112.0 112.0<br />

SW (º C) — 25.0 25.0 25.0 25.0 32.0 32.0 32.0<br />

CBS NO NO NO NO NO NO NO YES<br />

Most case studies corresponded to the limited operating conditions. The operating<br />

mode in each study was as follows:<br />

Case 1 The plant was only working as a full load power plant with no distilled<br />

water production (condensing mode).<br />

Case 2 The opposite <strong>of</strong> case 1. The plant was working as a pure distillation<br />

unit, producing only fresh water (desalination mode).<br />

Case 3 The nominal case: the plant was working at full load producing the<br />

maximum distilled water <strong>and</strong> maximum power (extraction mode).<br />

Case 4 The more usual operating conditions in winter (parallel mode).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 185


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

Case 5 Partial load operating conditions (extraction mode).<br />

Case 6 Minimum load operating conditions (parallel mode).<br />

Cases 7&8 The effect <strong>of</strong> the cleaning ball system was analyzed. In both cases some<br />

live steam was throttled in the reducing pressure station: the maximum<br />

load extracting live steam to a second MSF unit (twin extraction mode).<br />

The exergy <strong>and</strong> exergoeconomic costs <strong>of</strong> the most significant mass <strong>and</strong> energy<br />

*<br />

flowstreams (live steam generated in the boiler kBOI , steam to MSF vacuum system<br />

*<br />

*<br />

*<br />

kVST , steam to MSF brine heater kMSF , electric power kGEN <strong>and</strong> distilled water<br />

*<br />

kD ) appear in tables 7.10 <strong>and</strong> 7.11 respectively. No other energy <strong>analysis</strong> based on<br />

the First Law <strong>of</strong> Thermodynamics can provide this information, i.e., the amount<br />

(exergy or $) <strong>of</strong> the fuel plant consumed to obtain a flow.<br />

The unit costs in this section (the cost per unit exergy <strong>of</strong> the considered flow) only<br />

refer to the operating costs since they do not take into account the capital cost<br />

investment <strong>of</strong> the plant units.<br />

Afgan, Darwish <strong>and</strong> Carvalho (1999) quantified the primary energy or fuel needed to<br />

produce 1 kg <strong>of</strong> freshwater in a single purpose MSF desalination plant (case 2 in our<br />

<strong>analysis</strong>) <strong>and</strong> a dual purpose MSF desalination plant (case 3). These values (445 kJ<br />

<strong>and</strong> 225.7 kJ respectively) are based on an energy <strong>analysis</strong> <strong>of</strong> the dual-plant products<br />

<strong>and</strong> are quantitatively similar.<br />

Both tables provide the same information expressed in different units. The calculated<br />

costs are operating costs, discounting investment capital costs. The exergoeconomic<br />

costs c* were obtained by considering the natural gas market price (cf) <strong>of</strong> 2.35 ($/<br />

MBTU).<br />

TABLE 7.10 Exergy (kW fuel/kW product) unit costs k* <strong>of</strong> most significant flows <strong>of</strong> the dual plant.<br />

*<br />

kBOI *<br />

kVST *<br />

kMSF *<br />

kGEN k D *<br />

k * Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 Case 7 Case 8<br />

a<br />

2.733 2.371 2.604 2.590 2.576 2.572 2.677 2.559<br />

2.842 3.147 2.657 2.620 2.616 2.611 2.714 2.600<br />

— 3.871 2.693 2.644 2.667 2.650 3.650 3.615<br />

3.286 — 2.938 2.955 2.938 3.149 3.042 2.935<br />

— 416.511 221.67 224.99 227.67 261.02 549.48 526.38<br />

bD (kJ/kg) — 10.35 10.35 10.35 9.81 11.62 13.29 12.98<br />

a.<br />

*<br />

Exergy <strong>of</strong> water kD measured in a more realistic unit: (kJ fuel/kg water), therefore is included the exergy <strong>of</strong> water leaving the MSF<br />

unit (bD, in kJ/kg).<br />

186 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Cost <strong>analysis</strong><br />

TABLE 7.11 Exergoeconomic (monetary) unit costs ($/GJ) <strong>of</strong> most significant flows <strong>of</strong> a dual power <strong>and</strong><br />

desalination plant. Cost <strong>of</strong> water c* D is expressed in $/m 3 , <strong>and</strong> electricity cost <strong>of</strong> is also<br />

expressed in $/kW·h (c* GEN*).<br />

c* BOI<br />

c* VST<br />

c* MSF<br />

c* GEN<br />

c* *<br />

GEN<br />

c D *<br />

c * Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 Case 7 Case 8<br />

6.087 5.281 5.799 5.770 5.738 5.728 5.964 5.699<br />

6.331 7.010 5.913 5.835 5.827 5.816 6.046 5.792<br />

— 8.621 5.997 5.888 5.940 5.902 8.130 8.052<br />

7.319 — 6.543 6.583 6.545 7.014 6.775 6.537<br />

0.0263 — 0.0235 0.0237 0.0235 0.0252 0.0244 0.0235<br />

— 0.9277 0.4937 0.5011 0.5071 0.5814 1.223 1.172<br />

These values only contain the irreversibilities, the destruction <strong>of</strong> exergy or useful<br />

energy in the productive process. The live steam cost is always lower because it is<br />

generated at the very beginning <strong>of</strong> the production process. The irreversibilities during<br />

natural gas combustion <strong>and</strong> heat transfer inside the boiler increase the cost <strong>of</strong> this<br />

steam.<br />

Flowstreams further down the productive process were more costly. All processes in<br />

the plant were irreversible (see table 7.12) <strong>and</strong> the total exergy destroyed<br />

continuously increased throughout the productive process. The amount <strong>of</strong> exergy<br />

required to obtain a flow (exergy cost) also increased. For this reason, the final<br />

products had the highest costs.<br />

The effect <strong>of</strong> irreversibilities in the cost generation process is clearly shown by<br />

comparing studies 7 <strong>and</strong> 8. The cleaning ball system directly decreases distilled water<br />

cost by decreasing the irreversibility in the MSF plant (see table 7.12) <strong>and</strong> increasing<br />

efficiency (table 7.14). This benefit in the MSF plant also affects the power plant. The<br />

amount <strong>of</strong> steam needed in the MSF plant brine heater decreases (see table 7.9),<br />

increasing the steam mass flow rate exp<strong>and</strong>ed in the LP turbine <strong>and</strong> the electrical<br />

power produced. Modifying the operating conditions <strong>of</strong> the MSF affects the electrical<br />

cost.<br />

Irreversibilities (table 7.12) may have different costs. For example, boiler<br />

irreversibilities (I BOI) are much higher than MSF plant irreversibilities (I MSF), but live<br />

steam cost is lower than distilled water cost.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 187


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

TABLE 7.12 Irreversibilities (exergy destruction, kW) in the different components <strong>of</strong> the dual plant. MSF unit is<br />

considered a component.<br />

I Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 Case 7 Case 8<br />

I CP 89.26 58.44 13.93 16.53 13.84 4.94 7.94 16.52<br />

I LPH2 2,125.1 — 277.6 38.40 118.0 148.8 26.34 42.77<br />

I LPH1 1,687.8 — 363.0 80.85 279.0 268.2 28.97 219.7<br />

I DRT 306.0 655.5 990.4 544.1 1,362.4 638.7 821.1 2,317.7<br />

I FP 265.0 — 212.3 124.7 123.4 229.5 601.9 549.8<br />

I HPH2 243.2 — 505.4 274.1 329.7 77.61 454.3 566.6<br />

I HPH1 513.8 — 688.7 368.2 430.6 204.2 661.9 737.6<br />

I BOI 265,908.9 — 268,206.6 195,921.1 208,749.0 130,195.6 306,254.9 306,605.8<br />

I VST 89.62 — 1,157.2 394.8 493.7 105.0 423.0 442.7<br />

I HPT1 2,206.0 — 2,254.8 4,870.4 4,379.7 5,983.7 4,734.6 4,698.4<br />

I HPT2 250.5 — 637.2 451.1 489.8 112.8 467.1 479.8<br />

I HPT3 303.5 — 813.9 508.4 595.6 240.1 512.8 539.4<br />

I HPT4 955.4 — 2,369.5 1,097.1 1,464.0 583.0 912.5 1,054.3<br />

I LPT1 4,265.0 — 1,888.0 470.0 945.6 1,097.2 131.0 1,139.1<br />

I LPT2 9,598.9 — — — — 732.3 — 230.1<br />

I CND 37,816.7 — — — — 3,260.5 — 1,769.3<br />

I GEN 2,041.7 — — — — 1,372.9 — 1,489.5<br />

I VS1 — 38,198.1 — — — — 33,892.9 35,451.4<br />

I VS2 — 314.4 — — — — — —<br />

I VS3 — 1,750.2 — — — — — —<br />

I TOT-PP 329,734.2 185,880.0 292,042.4 206,923.9 226,061.2 145,255.9 351,501.1 358,339.3<br />

I TOT-MSF — 67,482.6 67,014.5 66,954.5 56,843.2 35,350.0 117,924.4 106,596.9<br />

The reasons for the impressive cost <strong>of</strong> distilled water are:<br />

• The large amount <strong>of</strong> exergy destruction (irreversibility) in the MSF plant,<br />

considering the high fuel value <strong>of</strong> the MSF unit (steam exhausted in brine heater<br />

<strong>and</strong> ejectors, electrical consumption) <strong>and</strong> the low value <strong>of</strong> the product<br />

(freshwater exergy flow). The energy <strong>and</strong> exergetic cost balance must be fulfilled<br />

(Valero, Muñoz <strong>and</strong> Lozano, 1986c).<br />

188 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Cost <strong>analysis</strong><br />

• The low distillate exergy flow is due to the low freshwater temperature leaving<br />

the MSF unit (see the last row in table 7.10. The different values stem from the<br />

different distilled water temperatures in different operating conditions (which<br />

strongly depends on the seawater temperature entering the desalination unit).<br />

The contribution <strong>of</strong> chemical <strong>and</strong> mechanical exergy to the global exergy flow <strong>of</strong><br />

seawater flows is minimum. Consequently, the final exergy cost is very low but<br />

the intermediate flows inside the distiller can be extremely high (the flashing<br />

brine, cooling brine, etc). The relationship between the inlet/outlet exergy flows<br />

which determine the exergy unit consumption k in the characteristic equations<br />

that model MSF thermoeconomics, is quite elevated in this example. The exergy<br />

unit consumption k propagates the exergy cost <strong>of</strong> the final product increasing the<br />

cost <strong>of</strong> water from the exergetic point <strong>of</strong> view.<br />

• The resources consumed in the MSF units are not primary energy. The electricity<br />

<strong>and</strong> steam produced to the distiller were produced in the power plant <strong>and</strong> the cost<br />

<strong>of</strong> the fuels <strong>of</strong> the MSF do not have a unit exergy cost. Only primary energy has<br />

an exergy cost equal to one (as natural gas entering the boiler).<br />

Another important result was the significantly higher water cost when the live steam<br />

was throttled through the HP reduction station (cases 2, 7 <strong>and</strong> 8). This has a physical<br />

explanation related with energy quality degradation. When the live steam exp<strong>and</strong>s<br />

through a throttle valve, its energy content remains stable while its exergy decreases<br />

(pressure is dramatically reduced in the reduction pressure station). The exergy<br />

destruction in the pressure reduction station correspond to I VS1, I VS2 <strong>and</strong> I VS3<br />

(table 7.12).<br />

Regarding component efficiencies, the more efficient a process the lower cost<br />

generated. Consider, for example, turbine efficiencies (table 7.13).<br />

TABLE 7.13 Isoentropic efficiencies <strong>of</strong> pumps <strong>and</strong> turbine sections <strong>of</strong> the power plant.<br />

η (%) Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 Case 7 Case 8<br />

HPT1 0.733 — 0.719 0.546 0.581 0.430 0.558 0.560<br />

HPT2 0.939 — 0.949 0.913 0.919 0.885 0.921 0.922<br />

HPT3 0.978 — 0.950 0.950 0.950 0.978 0.950 0.950<br />

HPT4 0.968 — 0.938 0.941 0.939 0.947 0.940 0.938<br />

HPT5 0.812 — 0.847 0.865 0.857 0.857 0.864 0.864<br />

LPT1 0.873 — 0.752 < 0 0.820 0.815 0.070 0.802<br />

LPT2 0.738 — 0.729 < 0 0.737 0.746 < 0 0.756<br />

FP 0.861 0.807 0.861 0.855 0.870 0.692 0.735 0.737<br />

CP 0.778 — 0.773 0.113 0.627 0.588 0.077 0.377<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 189


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

The live steam generated in the boiler is exp<strong>and</strong>ed in the HP turbines, before being<br />

extracted to the brine heater <strong>of</strong> the MSF plant. For this reason, the difference between<br />

the cost <strong>of</strong> steam to brine heater <strong>and</strong> the cost <strong>of</strong> live steam for the analyzed cases is<br />

directly related to the HP turbines efficiencies. Thus, the higher the HP turbine<br />

efficiency, the lower the cost difference in brine heater <strong>and</strong> live steam.<br />

A similar result is obtained for the cost difference between live steam <strong>and</strong> power<br />

generated. In the <strong>analysis</strong>, the low-pressure turbine efficiencies also influenced the<br />

observed differences.<br />

Table 7.14 contains global efficiency parameters for the whole plant <strong>and</strong> for the<br />

power <strong>and</strong> MSF plants. As in the device <strong>analysis</strong>, the more efficient the global<br />

process, the lower the cost <strong>of</strong> the final product. For example, in cases 7 <strong>and</strong> 8 the<br />

cleaning ball system clearly increases the exergy efficiency <strong>of</strong> the MSF plant <strong>and</strong> the<br />

whole plant. The distilled water <strong>and</strong> power cost decrease as a result. The exergetic<br />

efficiency we obtained for the MSF plant is similar to other estimate (Hamed<br />

et. al, 1999).<br />

TABLE 7.14 Product <strong>and</strong> fuel (kW), <strong>and</strong> exergetic efficiency (%) values for the power <strong>and</strong> MSF plants. Note:<br />

The efficiency <strong>of</strong> the boiler is not included in the final efficiency.<br />

Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 Case 7 Case 8<br />

P PP 144,260.4 64,712.9 186,024.8 140,244.6 144,096.3 87,113.3 191,289.9 184,994.0<br />

P MSF — 6,951.78 6,951.78 6,951.78 5,613.4 3,925.9 8,344.9 8,326.9<br />

F PP 473,994.6 250,592.9 478,067.2 347,168.5 370,157.5 229,369.2 542,791.0 543,333.3<br />

F MSF — 74,434.4 73,966.4 73,906.3 62,456.6 39,275.9 126,269.3 114,923.8<br />

η PP 30.4 0.0 38.9 40.4 38.9 36.7 35.2 34.0<br />

η MSF — 9.3 9.4 9.4 9.0 10.0 6.6 7.2<br />

η TOT 30.4 2.7 24.9 21.1 23.6 21.3 13.5 14.4<br />

Finally, product costs <strong>of</strong> different plant components were also calculated (see<br />

table 7.15).<br />

The steam leaving the boiler has a lower exergy cost since the fuel plant exergy only<br />

degraded in the boiler tubes (the combustion <strong>and</strong> heat transfer process is non-ideal).<br />

As the steam passes through the turbine section, its energy quality gradually<br />

degrades: the exergy cost increases from the first to the last turbine section. The<br />

exergy cost <strong>of</strong> the electricity is a weighted sum <strong>of</strong> the exergy costs <strong>of</strong> the turbine<br />

sections. The inefficiencies <strong>of</strong> the pumps increase the exergy cost <strong>of</strong> the electricity.<br />

190 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Cost <strong>analysis</strong><br />

The energy quality <strong>of</strong> the steam extracted for the heaters is degraded in this heating<br />

process. Although the live steam is the cheapest in desalination mode (case 2), the<br />

exergy cost <strong>of</strong> the steam to the MSF unit has a higher cost than the steam provided<br />

when the plant is producing electricity. The reduction pressure station is more<br />

inefficient than the set <strong>of</strong> components turbine-heaters-condenser.<br />

TABLE 7.15 Unit exergy costs k* (kW/kW) <strong>of</strong> component products in the steam power plant coupled with a<br />

MSF unit.<br />

k* Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 Case 7 Case 8<br />

*<br />

kCP 4.942 — 3.957 20.67 4.960 3.396 16.45 8.020<br />

*<br />

kLPH2 3.851 — 3.476 8.106 3.465 3.664 5.363 3.824<br />

*<br />

kLPH1 3.389 — 3.177 3.592 3.230 3.249 3.393 3.400<br />

*<br />

kDRT 3.040 4.306 3.021 2.944 3.283 3.280 3.059 3.473<br />

*<br />

kFP 3.764 3.151 3.282 3.272 3.223 4.077 3.771 3.560<br />

*<br />

kHPH2 2.992 — 2.850 2.793 2.796 2.729 2.926 2.822<br />

*<br />

kHPH1 3.034 — 2.853 2.796 2.797 2.789 2.930 2.813<br />

*<br />

kBOI 2.733 2.371 2.604 2.590 2.576 2.572 2.677 2.559<br />

*<br />

kVST 2.842 — 2.657 2.620 2.616 2.611 2.714 2.600<br />

*<br />

kHPT1 3.001 — 2.794 2.988 2.925 3.261 3.074 2.926<br />

*<br />

kHPT2 2.881 — 2.739 2.706 2.701 2.648 2.807 2.686<br />

*<br />

kHPT3 2.910 — 2.778 2.735 2.735 2.706 2.841 2.719<br />

*<br />

kHPT4 3.282 — 3.029 2.935 2.951 2.952 3.049 2.918<br />

*<br />

kLPT1 3.373 — 3.533 10.434 3.191 3.350 15.711 4.164<br />

*<br />

kLPT2 3.858 — 3.660 — 3.577 3.505 — 3.506<br />

*<br />

kVS1 — 3.851 — — — — — 4.591<br />

*<br />

kVS2 — 3.017 — — — — — —<br />

*<br />

kVS3 — 3.527 — — — — — —<br />

7.2.3 <strong>Thermoeconomic</strong> costs<br />

The thermoeconomic cost <strong>of</strong> a flow has two parts, one from the monetary cost <strong>of</strong> the<br />

fuel (natural gas) exergy needed to produce this flow, i.e., its exergoeconomic cost<br />

(Valero, Muñoz <strong>and</strong> Lozano, 1986b) <strong>and</strong> the other from the rest <strong>of</strong> the costs generated<br />

in the productive process (capital, maintenance, etc).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 191


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

The balance <strong>of</strong> thermoeconomic costs for any individual unit has one more term than<br />

the exergy cost balances (tables 7.7 <strong>and</strong> 7.8). The term Z/ϕ represents the contribution<br />

<strong>of</strong> the non-energetic production factors (investment capital costs). The balance <strong>of</strong><br />

thermoeconomic cost ($/s) is expressed in equation 7.5:<br />

c f F + Z/ϕ = cp P (7.5)<br />

where cf <strong>and</strong> cp are the unit thermoeconomic costs ($/kJ) <strong>of</strong> the fuel (F) <strong>and</strong> product<br />

(P) respectively. As the term Z is usually calculated in US dollars ($) it must be<br />

divided by a temporary factor, called amortization factor (ϕ). The amortization factor<br />

takes into account the economic life period <strong>of</strong> the plant <strong>and</strong> is also called the capital<br />

cost <strong>of</strong> an installation (see section 7.2.3.2 for more information on capital costs).<br />

7.2.3.1 Investment costs<br />

According to Bejan et al. (1997), an investment cost is a one-time cost, in contrast to<br />

fuel costs <strong>and</strong> O&M costs which are continuous or repetitive in nature. Investment<br />

costs are treated differently than fuel <strong>and</strong> O&M expenses in an economic <strong>analysis</strong>.<br />

Some concepts are necessary to underst<strong>and</strong> these costs:<br />

• Fixed capital investment, the total system capital cost assuming a zero-time<br />

design <strong>and</strong> construction period, i.e., the capital to purchase the l<strong>and</strong>, build all the<br />

necessary facilities <strong>and</strong> purchase <strong>and</strong> install the required machinery <strong>and</strong><br />

equipment.<br />

• Total capital investment, the sum <strong>of</strong> the fixed-capital investment <strong>and</strong> other<br />

outlays, i.e., startup costs, working capital, costs <strong>of</strong> licensing, research <strong>and</strong><br />

development, <strong>and</strong> allowance for funds used during construction.<br />

• Direct costs, the costs <strong>of</strong> all permanent equipment, materials, labor <strong>and</strong> other<br />

resources involved in the fabrication, erection, <strong>and</strong> installation <strong>of</strong> the permanent<br />

facilities.<br />

• Indirect costs, not a permanent part <strong>of</strong> the facilities but required for the orderly<br />

completion <strong>of</strong> the project: engineering <strong>and</strong> supervision, construction costs,<br />

contingencies. The fixed capital investment is the sum <strong>of</strong> direct <strong>and</strong> indirect costs.<br />

In our case, purchased-equipment costs provided by the plant managers are quite<br />

different from other studies (El-Sayed, 1996; Boehm, 1987; Frangopoulos, 1991;<br />

Lozano et al., 1996). This is due to the magnitude <strong>of</strong> the components considered in<br />

the dual plant. Several authors propose costing equations for most <strong>of</strong> the components<br />

used in our <strong>analysis</strong>, but the main parameters used in the proposed correlations are<br />

outside the specified range (our power plant <strong>and</strong> desalination units were very large).<br />

Other costs not included in the capital costs <strong>of</strong> the components (but that also<br />

constitute a part <strong>of</strong> the direct costs <strong>of</strong> the fixed-capital investment) include the<br />

purchased-equipment installation, piping, instrumentation <strong>and</strong> controls, electrical<br />

equipment <strong>and</strong> materials, l<strong>and</strong>, civil <strong>and</strong> structural work <strong>and</strong> service facilities.<br />

192 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Cost <strong>analysis</strong><br />

El-Sayed (1996) calculates the cost (in thous<strong>and</strong>s <strong>of</strong> dollars, k$) <strong>of</strong> the main<br />

components <strong>of</strong> the equipment used in a MSF <strong>and</strong> steam power plant using the<br />

following equation:<br />

Z = ca A, (7.6)<br />

where the area A is calculated using an exponential formula as a function <strong>of</strong> four<br />

parameters, i.e.:<br />

A =<br />

k x1 n1 n2 n3 n4<br />

x2 x3 x4<br />

These parameters are shown in table 7.16.<br />

TABLE 7.16 Costing equation parameters for an MSF <strong>and</strong> power plant (El-Sayed, 1996). Units: ca k$/ft 2 ,<br />

A ft 2 , M lb/s, Q kW, P i , P e psia, T i R, ∆T F, ∆P, dP psi, e = η/1– η. Subscripts: i, inlet; e, exit; t,<br />

tube; s, shell; m, mean (LTMD).<br />

Component ca k x 1 x 2 x 3 x 4 n1 n2 n3 n4<br />

Steam turbine 50 0.45 M T i /P i P e e 1 0.05 –0.75 0.9<br />

Feed pump 3 0.0025 M ∆P e — 1 0.55 1.05 —<br />

C.W. pump 3 0.0063 M ∆P e — 1 0.1 0.7 —<br />

Economizer 0.015 310 Q ∆T m dP t dP s 1 –1 –0.16 –0.12<br />

Boiler 0.015 340 Q ∆T m dP t dP s 1 –1 –0.33 –0.26<br />

Superheater 0.015 310 Q ∆T m dP t dP s 1 –1 –0.15 –0.14<br />

Heater 0.02 3.3 Q ∆T t dP t dP s 1 –0.7 –0.08 –0.04<br />

MSF 0.02 10 Q ∆T n ∆T t dP t 1 –0.75 –0.5 –0.1<br />

(7.7)<br />

Boehm (1987) introduces the size effect <strong>of</strong> the units into a simple cost equation that<br />

only depends on a variable, S. Thus, a complete tabulation <strong>of</strong> data for a particular<br />

piece <strong>of</strong> equipment could contain reference cost <strong>and</strong> size (Z r <strong>and</strong> S r), <strong>and</strong> the factor m<br />

responsible for the economies <strong>of</strong> scale.<br />

Z = Z r (S/S r) m (7.8)<br />

In the cost equation, Boehm normally uses a range <strong>of</strong> 0.5-1.0. Sometimes m is greater<br />

than 1.0 (boilers, heaters…), which produces unexpected results. Table 7.17 includes<br />

the main parameters <strong>of</strong> the above mentioned equation.<br />

Finally, the more accurate equations, in comparison with the cost estimation provided<br />

by the plant managers, are those proposed by Frangopoulos (1991). They are usually<br />

a correlation with three or four main parameters <strong>and</strong> correction factors depending on<br />

the device (see table 7.18).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 193


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

TABLE 7.17 Component parameters in Boehm (1987) equations.<br />

Component Z r S S r m<br />

Pump 47 M 10 0.03<br />

Steam turbine 25 W 1000 0.68<br />

Heater 21 A 100 0.71<br />

Condenser 3 Q 10 0.55<br />

Boiler 340 M 12 0.67<br />

TABLE 7.18 Costing equations proposed by Frangopoulos (1991).<br />

Component Cost equation<br />

Boiler 20.1552224 * exp (0.0014110546 * P1 ) * exp (0.7718795 * ln (M1 )) * FAR * FAN * FAT<br />

Steam Turbine 5240.378 * exp (0.569323 * ln (FB1 * (F2T + F2P))) * FBN * FBT<br />

Condenser 1.11 * A * 426.2632633 * exp (–0.4556513 * ln (A)) * FCR * FCPW * FCP * FCB<br />

Pump 1969.2325 * exp (0.4838546 * ln (7.279088e – 5 * M1 * 0.018 * (P2 –P1 ) * FDN<br />

Heater a<br />

Exp (8.202 + 0.01506 * ln (A) + 0.06811 * (ln (A)) 2 ) * FD * FP * FM<br />

Factor Correction factor<br />

FAR FAR = 1.0 + ((1–∆P r )/(1–∆P)) 8<br />

FAN FAN = 1.0 + ((1 – η1 r )/(1– η1)) 7<br />

FAT FAT = 1.0 + 5 * exp ((T1 – 1100)/18.75)<br />

FB1 FB1 = 0.0003929119 * η * M1 F2T F2T = 0.55 * (T1 – T2 – T2 * ln (T1 /T2 ))<br />

F2P F2P = 0.1102109 * T2 * ln (P1 /P2 )<br />

FBN FBN = 1 + ((1 – η r )/(1 – η)) 3<br />

FBT FBT = 1.0 + 5 * exp ((T 1 – 1100)/18.75)<br />

FCR FCR = (P 3 * ((1/∆P s ) – 1)/14.7) –0.11<br />

FCPW FCPW = (∆P t /14.7) –0.38<br />

FCP FCP = 0.93 + 2.6380952 e –4 * P2 + 1.352381 e –6 2<br />

* P2 FCB FCB = exp (0.10/(TTD–5))<br />

FDN FDN = 1 + ((1 – 0.8)/(1 – η)) 3<br />

FD FD = exp (–0.7844 + 0.083 * LN (A))<br />

FP FP = 0.8955 + 0.04981 * LN (A)<br />

FM FM = 1.4144 + 0.23296 * LN (A)<br />

a. From Chemical Engineering (Corripio, Chrien <strong>and</strong> Evans, 1982). P 1 , T 1 <strong>and</strong> M 1 are the inlet conditions, T 2 , P 2 the exit conditions,<br />

A area, η <strong>and</strong> η1 efficiency <strong>and</strong> First principle efficiency, TTD terminal temperature difference, ∆P s , ∆P t pressure losses in tubes <strong>and</strong><br />

shell.<br />

194 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Cost <strong>analysis</strong><br />

Lozano et al. (1996) also propose a set <strong>of</strong> equations for a wide range <strong>of</strong> values to<br />

obtain a reasonable equipment cost (see table 7.19):<br />

TABLE 7.19 Cost equations proposed by Lozano et al. (1996). η exergetic efficiency, B exergy flow <strong>of</strong><br />

product, S negentropy, vw velocity <strong>of</strong> tubes , W power, e eficiency <strong>of</strong> the condenser (= T 0 (s 2 –s 1 )/<br />

(h 2 –h 1 )).<br />

Component Cost equation<br />

Boiler 740 * exp ((P 1 –28)/150) * (1 + 5 * exp ((T 1 –866)/10.42)) * (1 + ((0.45–0.405)/(0.45–η)) 7 ) * B 0.8<br />

St. Turbine 3000 * (1 + 5 * exp ((T 1 –866)/10.42)) * (1 + ((1–0.953)/(1–η)) 3 ) * W 0.7<br />

Condenser (1/(T 0 * e)) (217 * (0.247 + 1/(3.24 * vw 0.8 )) * ln (1/(1–e)) + 138) * (1/(1–η)) * S<br />

Pump 378 * (1 + ((1–0.808)/(1–η)) 3 ) * B 0.71<br />

Purchase cost provided by the plant managers is much more complete than the<br />

individual components. It includes the price breakdown per section <strong>of</strong> each unit, <strong>and</strong><br />

the direct costs <strong>of</strong> the installation. Table 7.20 includes a list with the percentages <strong>of</strong><br />

each unit or subsystem with respect the total purchase cost (direct cost) <strong>of</strong> a power<br />

<strong>and</strong> desalination plant. L<strong>and</strong> cost is neglected in the Gulf Area.<br />

The price breakdown in table 7.20 does not contain the cost <strong>of</strong> each component in the<br />

productive structure. As a result, the thermoeconomic cost can only be calculated for<br />

the final products in the power <strong>and</strong> desalination plant, knowing the exergy cost <strong>of</strong> the<br />

electricity <strong>and</strong> distillate, the economic investment cost <strong>and</strong> the thermoeconomic cost<br />

<strong>of</strong> the products. The thermoeconomic cost can be expressed in units <strong>of</strong> money per<br />

unit <strong>of</strong> time ($/s), or units <strong>of</strong> money per unit <strong>of</strong> product: $/kW·h or $/m 3 . All cost data<br />

must have the same reference year as a basis for calculations. This is done with an<br />

appropriate cost index, an inflation indicator from technical journals (e.g. Chemical<br />

Engineering) that corrects the cost <strong>of</strong> equipment. We did not apply the cost index<br />

since the purchase costs <strong>of</strong> our installation were updated in 1997.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 195


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

TABLE 7.20 Price breakdown per section in a dual-purpose plant.<br />

Component system Portion<br />

Steam Turbine Plant 12,25<br />

HP heater, LP heater, feedwater storage tank with deaerator, cold condensate<br />

storage tank<br />

1,13<br />

Steam generating plant 13,15<br />

HP feeding system 0,12<br />

LP feeding system 0,30<br />

Boiling feed pump sets with hydraulic coupling 1,66<br />

Generator complete with air cooling <strong>and</strong> excitation systems 2,63<br />

Others: Transformers, busbars, switchboards, cabling <strong>and</strong> cable laying,<br />

rectifiers, batteries, electrical control equipment, instrumentation <strong>and</strong> control,<br />

service water <strong>and</strong> drainage system.<br />

196 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

14,08<br />

Total for the steam power plant 45,32<br />

MSF unit: Evaporator shell <strong>and</strong> tube bundles 20,38<br />

Brine heater 1,06<br />

Deaerator 0,04<br />

Vacuum system 0,63<br />

Cooling water recircul. pump set including isolating, non-return valves 0,29<br />

2 Brine recirculating pump sets, complete 0,87<br />

Blow down pump set, complete 0,22<br />

2 Distillate pump units 0,22<br />

2 Brine heater condensate pump sets, complete 0,07<br />

Others: Protective coating, make-up water strainers, cranes, seawater, brine<br />

recirculation, blowdown <strong>and</strong> distillate pipeline, HP, MP <strong>and</strong> LP reducing<br />

stations, antiscaling, antifoaming <strong>and</strong> sodiumsulfite systems, on-load tube<br />

cleaning system, lighting system, instrumentation <strong>and</strong> control, switchgear,<br />

switchboards, transformer.<br />

Total for the desalination unit 32,79<br />

General services: Circulating water <strong>and</strong> seawater supply system, seawater<br />

cleaning plant, fuel oil <strong>and</strong> gas system, power transformers, bus duct systems,<br />

cables, lighting <strong>and</strong> power outlets, earthing system, common instrumentation<br />

<strong>and</strong> control, water treatment, lifts, buildings, fire fighting systems, chemicals<br />

<strong>and</strong> chlorination system, town water storage, DPS system, chemical storage.<br />

9,01<br />

21,89<br />

Total for the dual plant 100


Cost <strong>analysis</strong><br />

7.2.3.2 Capital costs<br />

The average capital cost for the system was assumed to be 3.47×10 –9 $/s·$. It was<br />

calculated based on 8% capital recovery per calendar year (8,000 hours operation a<br />

year) <strong>and</strong> 15% allowance for the fixed part <strong>of</strong> O&M (El-Sayed, 1996). The average<br />

capital cost takes into account the effect <strong>of</strong> inflation: price increases associated with<br />

increase in available currency <strong>and</strong> credit without a proportional increase in available<br />

goods <strong>and</strong> services <strong>of</strong> the same quality. This cost also includes the effect <strong>of</strong> escalation<br />

(resource depletion, increased dem<strong>and</strong> <strong>and</strong> technological advances); <strong>and</strong> depreciation<br />

(decrease in equipment value due to physical deterioration, technological advances<br />

<strong>and</strong> replacement). Some assumptions were made to assess the average capital cost.<br />

For example, l<strong>and</strong> costs <strong>and</strong> total capital investment were placed at the beginning <strong>of</strong><br />

the design <strong>and</strong> construction period so that the end <strong>of</strong> this period is considered the<br />

beginning <strong>of</strong> commercial operation (economic-life period).<br />

7.2.4 <strong>Thermoeconomic</strong> cost <strong>analysis</strong><br />

The exergy <strong>and</strong> economic costs <strong>of</strong> a system provide the real plant operating costs.<br />

Tables 7.21 <strong>and</strong> 7.22 show the thermoeconomic cost in the eight cases (see table 7.9<br />

for details).<br />

TABLE 7.21 <strong>Thermoeconomic</strong> costs <strong>of</strong> distilled water <strong>and</strong> electricity <strong>of</strong> the analyzed dual-purpose plant.<br />

Cost ($/s) Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 Case 7 Case 8<br />

Electricity 1.5798 0.5068 1.3046 1.0030 1.1019 0.8818 1.0248 0.9706<br />

Water 0.3571 0.9798 0.6885 0.6935 0.6471 0.5534 1.1251 1.1088<br />

TABLE 7.22 <strong>Thermoeconomic</strong> cost <strong>of</strong> electricity ($/kW·h) <strong>and</strong> water ($/m 3 ) for the cases studied in the<br />

exergetic cost <strong>analysis</strong>.<br />

Cost Case 1 Case 2 Case 3 Case 4 Case 5 Case 6 Case 7 Case 8<br />

Electricity 0.0388 0 0.0385 0.0479 0.0436 0.0593 0.0482 0.0492<br />

Water 0 1.5026 1.0558 1.0635 1.1648 1.6871 1.8456 1.7802<br />

El-Sayed (1996) proposes the following costs for the products <strong>of</strong> a typical dualpurpose<br />

power <strong>and</strong> desalination plant:<br />

• Electricity: 0.045 $/kW·h.<br />

• Water: 1.3 $/m 3 .<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 197


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

The results <strong>of</strong> the thermoeconomic <strong>analysis</strong> were very close to the values given by El-<br />

Sayed, especially in the most representative cases (in hours <strong>of</strong> operation per year,<br />

cases 3, 4 <strong>and</strong> 5). Note that the thermoeconomic cost was not zero in case 1 for water<br />

nor for electricity in case 2 (see table 7.21) despite the lack <strong>of</strong> production. This was<br />

due to the effect <strong>of</strong> amortization <strong>of</strong> the purchase costs in the first table. The effect on<br />

quantity production is clear in table 7.22 (the cost <strong>of</strong> electricity per unit <strong>of</strong> energy is<br />

reduced in case 1 <strong>and</strong> is lower than other costs, although this is the worst case if we<br />

analyze exergy costs). In Case 6 (with partial load) the investment costs overcharge<br />

the cost per unit <strong>of</strong> production. The use <strong>of</strong> the reduction pressure station to produce<br />

freshwater is not recommended even with a high freshwater dem<strong>and</strong> (see cases 2, 7<br />

<strong>and</strong> 8 in table 7.22). Case 3 is the most interesting case to maintain the best operation<br />

mode.<br />

7.2.5 Cost allocation: Indirect methods<br />

Some cost allocation methods allocate the total cost <strong>of</strong> owning <strong>and</strong> operating the<br />

plant among two products, without having to split the total cost in two products<br />

(direct methods). Other methods allocate the main factory costs (e.g. manpower,<br />

material, fuel <strong>and</strong> capital depreciation) among the two products (indirect methods).<br />

Some criterion is usually needed to help in cost allocation. For example, the exergy<br />

cost method is an indirect method that allocates the cost <strong>of</strong> producing the two<br />

products in terms <strong>of</strong> fuel consumption.<br />

Although cost allocation methods are a rational basis for pricing the two products, the<br />

cost is the amount <strong>of</strong> resources needed to obtain these products. The price imposed on<br />

a product is independent <strong>of</strong> the efficiency <strong>of</strong> the formation process <strong>of</strong> that product.<br />

7.2.5.1 WEA method<br />

The method proposed by El-Nashar (1999) <strong>and</strong> the Water <strong>and</strong> Electricity Department<br />

<strong>of</strong> the UAE (WEA method) is indirect <strong>and</strong> allocates all cost components among water<br />

<strong>and</strong> electricity according to functional considerations. The annual cost for a cogeneration<br />

plant can usually be separated into three cost components: fixed capital<br />

charges, fuel costs <strong>and</strong> O&M costs. Each one can be separated into costs for<br />

electricity production, costs for heat production <strong>and</strong> common costs to both products.<br />

The methods differ in how they separate annual costs into the three components <strong>and</strong><br />

in allocating common costs between electricity <strong>and</strong> heat.<br />

The total costs are divided into five cost departments: fuel, personnel, maintenance<br />

contracts, spares <strong>and</strong> consumables <strong>and</strong> depreciation <strong>of</strong> fixed capital. Personnel costs<br />

are divided among those directly involved in the co-generation plant (such as<br />

operation <strong>and</strong> maintenance work), or those that serve several plants. The cost <strong>of</strong> fuel<br />

consumed by the steam turbines is split between electricity <strong>and</strong> water since the steam<br />

derived to the MSF unit has the potential to generate a certain amount <strong>of</strong> electrical<br />

198 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Cost <strong>analysis</strong><br />

power if allowed to exp<strong>and</strong> through a hypothetical condensing turbine. Since this<br />

steam is used for desalination instead <strong>of</strong> additional power generation, the fuel<br />

consumed for this amount <strong>of</strong> non-produced electrical power should be charged to<br />

water. The amount <strong>of</strong> additional power (W CT) which could have been generated by<br />

this hypothetical turbine (in our case is the low pressure turbine) may be expressed<br />

as:<br />

W CT = Q η B η CT<br />

(7.9)<br />

where Q is the amount <strong>of</strong> heat supplied to the hypothetical steam turbine, η B is the<br />

efficiency <strong>of</strong> boiler <strong>and</strong> η CT is the thermal efficiency <strong>of</strong> the condensing steam turbine<br />

cycle. The fuel consumption Gc could be allocated to electricity <strong>and</strong> water according<br />

to the following equations, taking into account the power generated in the real steam<br />

turbine (W ST):<br />

Gc e = Gc W ST /(W ST + W CT) (7.10)<br />

Gc w = Gc W CT /(W CT + W CT) (7.11)<br />

The fuel allocation problem could also be solved using the difference in output power<br />

produced when the amount <strong>of</strong> fuel consumed is the same in both cases. The MR (no<br />

desalination) <strong>and</strong> MCR (co-generation) cases are a good example. The total<br />

personnel cost consists <strong>of</strong> directly assessable costs (e.g. operating <strong>and</strong> maintenance<br />

staff) <strong>and</strong> indirect or common service personnel. The directly assessable portions are<br />

charged to either electricity or water, depending on the case. The cost <strong>of</strong> common<br />

service personnel is allocated to electricity <strong>and</strong> water according to the ratio <strong>of</strong> the<br />

capital cost <strong>of</strong> the plant <strong>and</strong> equipment associated with electricity production <strong>and</strong><br />

desalination. Maintenance contracts for specialized maintenance work is priced <strong>and</strong><br />

electricity <strong>and</strong> water are finally allocated. Depreciation <strong>of</strong> capital cost between<br />

electricity <strong>and</strong> water is allocated according to the function <strong>of</strong> the equipment in<br />

operation. The depreciation cost is allocated to electricity in the steam turbine power<br />

plant <strong>and</strong> water in the desalination plant. Depreciation costs <strong>of</strong> common equipment<br />

<strong>and</strong> facilities are allocated according to the capital cost <strong>of</strong> equipment related to<br />

electricity <strong>and</strong> water, as done for the common personnel costs.<br />

The WEA method is widely used in the UAE to allocate the cost <strong>of</strong> producing water<br />

<strong>and</strong> electricity in co-generation plants (starting from the yearly electrical <strong>and</strong> water<br />

production) <strong>and</strong> the cumulative number <strong>of</strong> operating hours <strong>of</strong> power <strong>and</strong> desalination<br />

plants. Those data are confidential <strong>and</strong> cannot be presented here. Other characteristics<br />

include:<br />

• Average yearly cost (with a wide range <strong>of</strong> operating modes) <strong>of</strong> the co-generation<br />

plants (with several configurations <strong>of</strong> dual plants) operating in the country. It is<br />

not valid for calculating an instantaneous cost <strong>of</strong> water <strong>and</strong> electricity.<br />

• Applied fuel <strong>and</strong> capital costs (which are unknown).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 199


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

Compared with exergy cost methodology, the trend <strong>of</strong> water <strong>and</strong> electricity costs is<br />

the following:<br />

• The WEA method tends to overvalue electricity <strong>and</strong> undervalue water by<br />

charging all the capital <strong>and</strong> O&M costs <strong>of</strong> the steam turbine (except fuel) to<br />

electricity.<br />

• The WEA costing methodology only allocates fuel cost to steam turbines.<br />

• The WEA method suffers from a certain degree <strong>of</strong> arbitrariness with regard to<br />

the efficiency <strong>of</strong> a hypothetical condensing steam turbine. The assumptions<br />

could cause fluctuations in the resulting cost <strong>of</strong> electricity <strong>and</strong> water. The<br />

difference in production between operating modes could partially avoid this<br />

problem (see next section).<br />

• The exergy/thermoeconomic method charges each product <strong>of</strong> a multi-product<br />

unit to the appropriate portion <strong>of</strong> capital <strong>and</strong> O&M costs involved in operating<br />

the unit.<br />

• The exergy/thermoeconomic method is based on a solid accounting <strong>and</strong><br />

thermodynamics. Therefore, it will be used in our studies.<br />

As a result <strong>of</strong> the above, El-Nashar (1993; 1999) developed a model based on exergy<br />

<strong>analysis</strong> to predict the final costs <strong>of</strong> the two products. Other authors propose cost<br />

redistribution using the exergy <strong>analysis</strong> <strong>of</strong> the dual-purpose plant (Evans, Crellin <strong>and</strong><br />

Tribus, 1980; Breidenbach, Rautenbach <strong>and</strong> Tusel, 1997; Slesarenko <strong>and</strong> Shtim,<br />

1986). The energy efficiency <strong>of</strong> the dual-purpose plant is also used to allocate the<br />

fuels to power <strong>and</strong> desalination <strong>and</strong> the relevant specific fuel costs for power<br />

generation <strong>and</strong> water production (Saeed, 1992).<br />

7.2.5.2 Fuel cost <strong>of</strong> water in dual plants<br />

Fuel energy for desalting depends on fuel allocation rules between the power <strong>and</strong><br />

desalted water produced in a dual-purpose plant (Darwish, Yousef <strong>and</strong> Al-Najem,<br />

1997). Kronenberg <strong>and</strong> Dvornikov (1999) argues that the steam cost <strong>of</strong> desalting<br />

should be calculated by defining the heat rate difference between the power plant<br />

coupled <strong>and</strong> uncoupled to the desalination plant (also called the Lost Kilowatts<br />

Method, see Gaggioli <strong>and</strong> El-Sayed, 1987; El-Saie <strong>and</strong> El-Saie, 1989). This heat rate<br />

difference is defined by the Fuel Cost <strong>of</strong> Water (FCW) in a dual-purpose installation.<br />

The fuel cost <strong>of</strong> water largely depends on the overall efficiency <strong>of</strong> the power plant,<br />

<strong>and</strong> is calculated as:<br />

FCW $ m 3<br />

( ⁄ )<br />

( W1 – W2) HR1 cf<br />

=<br />

----------------------------------------------<br />

Qf D<br />

200 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(7.12)


Cost <strong>analysis</strong><br />

where W 1 <strong>and</strong> W 2 are the electric power output <strong>of</strong> the uncoupled <strong>and</strong> coupled plant<br />

(kW), HR 1 is the heat rate <strong>of</strong> the uncoupled power plant (the inverse <strong>of</strong> the efficiency,<br />

kJ/kW·h), cf is the fuel cost ($/kg), Qf is the heat value <strong>of</strong> fuel (kJ/kg) <strong>and</strong> D is the<br />

water production (m 3 /h).<br />

Fuel cost <strong>of</strong> water can be calculated in the dual-purpose plant. For instance, the FCW<br />

<strong>of</strong> the MCR case was calculated using natural gas with a high heating value<br />

(HHV = 9,500 kcal/m 3 ), a density <strong>of</strong> 0.75 kg/m 3 <strong>and</strong> an energy cost <strong>of</strong> 2.23×10 –6 $/kJ<br />

(applied in the cost <strong>analysis</strong>). The gas consumption in the MR case (the uncoupled<br />

power plant in our case) was 43,500 Nm 3 /h. The final values to be introduced in<br />

formula (7.12) for our example are also introduced after the FCW value:<br />

FCW = 0.271 $/m 3<br />

W 1 = 146,700 kW<br />

W 2 = 122,000 kW<br />

HR 1 = 43,500 · (9,500 · 4.1868)/146,700 = 11,794 ·1 kJ/kW·h<br />

cf = 2.23 ·10 –6 (9,500 · 4.1868)/0.75 = 0.1182 $/kg<br />

Q = (9,500 · 4.1868)/0.75 = 53,032.8 kJ/kg<br />

D = 2,400 m 3 /h<br />

Note that the exergy <strong>analysis</strong> <strong>and</strong> the lost kilowatts method are similar (see section<br />

7.1.1 for the exergy <strong>analysis</strong> <strong>of</strong> the simple co-generation plant), although the latter<br />

uses the energy <strong>analysis</strong> to calculate the cost <strong>of</strong> fuel consumed in the co-generation<br />

plant. The resulting cost <strong>of</strong> water is very similar in both methods.<br />

If the FCW is compared with the exergoeconomic cost <strong>of</strong> case 3 in table 7.11 (i.e., the<br />

exergoeconomic cost <strong>of</strong> the MCR case), the FCW is more or less 55% <strong>of</strong> the<br />

thermoeconomic cost (0.493 $/m 3 ). The difference is mainly due to several factors:<br />

The exergoeconomic cost also includes the cost <strong>of</strong> electricity needed to pump the<br />

MSF flows <strong>and</strong> the steam derived to the vacuum system <strong>of</strong> the distillers.<br />

The FCW assumes a constant efficiency in the power plant (the heat rate <strong>of</strong> the plant<br />

in condensing mode). The overall efficiency <strong>of</strong> the dual-purpose plant is lower when<br />

the plant is only generating electricity (see table 7.14 for the exergetic efficiency <strong>of</strong><br />

the whole plant). Therefore, the amount <strong>of</strong> additional electricity generated in the<br />

condensing mode is not a valid index to calculate the fuel cost in co-generation mode,<br />

with a higher efficiency.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 201


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

7.3 <strong>Thermoeconomic</strong> diagnosis<br />

Diagnosis is the identification <strong>of</strong> something that is not working properly.<br />

<strong>Thermoeconomic</strong> diagnosis is the only operation <strong>analysis</strong> based on the Second Law.<br />

It uses the exergy balance <strong>of</strong> an installation to allocate <strong>and</strong> calculate irreversibilities<br />

in the production process <strong>and</strong> identify the equipment affecting overall efficiency. In<br />

practice, however, this useful information is not sufficient since some irreversibilities<br />

cannot be avoided. The technical possibilities for saving energy are always lower than<br />

the theoretical limit <strong>of</strong> thermodynamic energy losses. Moreover, the local exergy<br />

savings in different units or processes are not equivalent. The same local<br />

irreversibility decrease in two different components generally produces different<br />

variations in the total energy consumption.<br />

The final objective <strong>of</strong> <strong>Thermoeconomic</strong> diagnosis is to describe how malfunctions<br />

affect additional resource consumption (see Chapter 6 for a review <strong>of</strong><br />

<strong>Thermoeconomic</strong> theory <strong>and</strong> its applications). In this section, we analyze a power <strong>and</strong><br />

desalination plant according to the principles outlined in the previous chapter. The<br />

entire diagnosis is presented using the Structural Theory <strong>of</strong> <strong>Thermoeconomic</strong>s<br />

(Valero et al., 1993). It provides information about component fuel consumption<br />

during equipment degradation (inefficiency), how each component increases fuel<br />

consumption <strong>and</strong> how a component's inefficiency affects the behavior <strong>of</strong> other plant<br />

units.<br />

We will only consider the direct problem <strong>of</strong> thermoeconomic diagnosis (Valero,<br />

Torres <strong>and</strong> Lerch, 1999), where inefficiencies are quantified in terms <strong>of</strong> irreversibility<br />

increase, while distinguishing between efficiency deterioration (intrinsic <strong>and</strong> induced<br />

malfunctions) <strong>and</strong> component dysfunction (generated by the malfunction). The<br />

inefficiencies were previously simulated <strong>and</strong> the causes <strong>of</strong> the behavior deviation<br />

provoked by this inefficiency are not searched here.<br />

The inverse problem is to identify <strong>and</strong> quantify malfunctions (the origin <strong>of</strong> new<br />

irreversibilities). Classical thermoeconomic <strong>analysis</strong> does not elucidate the cause <strong>of</strong><br />

irreversibilities, although an effort is made to detect <strong>and</strong> stop malfunctions. The<br />

inverse problem finds the cause <strong>of</strong> the deviation between two states <strong>of</strong> the plant<br />

(actual <strong>and</strong> reference conditions). It requires a data acquisition system (for the<br />

reference conditions), a simulator (to provide the reference state for the same<br />

operating conditions) <strong>and</strong> conventional methods <strong>of</strong> the thermoeconomic diagnosis<br />

(the direct problem). One <strong>of</strong> the main difficulties with the inverse problem is<br />

recognizing <strong>and</strong> separating effects not intimately related with the inefficiencies <strong>of</strong> the<br />

plant components, such as load variation, set points or ambient conditions.<br />

The impact on fuel predicted by the simulator is exactly the same as that calculated<br />

by the Structural Theory <strong>of</strong> <strong>Thermoeconomic</strong>s. This plant diagnosis reproduces the<br />

deviation <strong>of</strong> the physical values when one or more inefficiencies are detected.<br />

202 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> diagnosis<br />

Although the simulator calculates the thermodynamic state <strong>of</strong> the dual plant with<br />

reasonable accuracy under different operating conditions, it might not be able to<br />

respond as well to unexpected non-linear inefficiencies. The diagnosis involves a<br />

sensitivity <strong>analysis</strong> <strong>of</strong> the mathematical model <strong>of</strong> the dual-purpose plant (simulator)<br />

with respect to a parameter (in this case, one or several inefficiencies in a component<br />

<strong>of</strong> the system). The simulator could be avoided in the diagnosis if the data acquisition<br />

system <strong>of</strong> the dual plant were available.<br />

We will first summarize the different inefficiencies, loads <strong>and</strong> operating modes<br />

simulated in the plant diagnosis. Then, the ‘direct problem’ <strong>of</strong> diagnosing one or<br />

several inefficiencies is analyzed for a defined load (corresponding to an operating<br />

mode) in the power <strong>and</strong>/or desalination plant. The <strong>analysis</strong> involves a new technique<br />

(see Chapter 6, Torres et al., 1999) based on Structural Theory <strong>and</strong> Symbolic<br />

<strong>Thermoeconomic</strong>s to provide a huge quantity <strong>of</strong> information, including:<br />

1. The irreversibility generated in each component.<br />

2. The exergetic cost <strong>of</strong> each component's product.<br />

3. The intrinsic malfunction in each component (i.e. the efficiency decrease <strong>of</strong> a<br />

component due to its own inefficiency).<br />

4. The induced malfunction in each component (i.e. the efficiency decrease <strong>of</strong> a<br />

component due to inefficiencies in other components).<br />

5. The dysfunction induced in the component due to the malfunction or<br />

inefficiency <strong>of</strong> other subsystems, which forces it to consume more local<br />

resources to attain the additional production required by the other components.<br />

6. The fuel impact or malfunction cost <strong>of</strong> each component due to an inefficiency,<br />

<strong>and</strong> the total impact on fuel.<br />

7. A compact <strong>and</strong> easy to underst<strong>and</strong> malfunction matrix containing the cost <strong>of</strong><br />

inefficiencies <strong>and</strong> the effect <strong>of</strong> a component inefficiency on all other<br />

components.<br />

7.3.1 <strong>Thermoeconomic</strong> diagnosis <strong>of</strong> a power <strong>and</strong> desalination<br />

plant: case studies<br />

System operating parameters can be classified according to their effect on component<br />

efficiency:<br />

• Local variables, which mainly affect the behavior <strong>of</strong> the component related to<br />

the variable (e.g. the isoentropic efficiency <strong>of</strong> a turbine).<br />

• Global or zonal variables, where the operating parameter cannot be associated<br />

with a specific component (e.g. live steam conditions <strong>of</strong> a steam power plant).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 203


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

A variable is considered local if the total impact on fuel associated with a subsystem<br />

is basically located in this component.<br />

We simulated the device inefficiencies <strong>and</strong> considered the different <strong>simulation</strong> data as<br />

plant data under different conditions (including inefficiencies). All the analyzed<br />

inefficiencies were associated with local plant variables <strong>and</strong> were chosen in terms <strong>of</strong><br />

their effect on energy:<br />

• Degradation <strong>of</strong> the isoentropic efficiency <strong>of</strong> the high-pressure turbine (1st section,<br />

HPT1, <strong>and</strong> 4th section HPT4).<br />

• Degradation <strong>of</strong> the isoentropic efficiency <strong>of</strong> the low-pressure turbine (1st section,<br />

LPT1).<br />

• Heat transfer problems in HP heaters were analyzed by varying the Terminal<br />

Temperature Difference TTD (temperature difference between the saturation<br />

temperature <strong>of</strong> the steam extracted from the turbine <strong>and</strong> feedwater leaving the<br />

heater). Only the HP heater no. 1 (HPH1) was treated.<br />

• By varying the feed pump isoentropic efficiency, operating inefficiencies were<br />

simulated in the feed pump.<br />

The effect <strong>of</strong> a global variable such as live steam temperature can be studied if the<br />

simulator supports a non-fixed condition in the live steam leaving the boiler. In the<br />

case <strong>of</strong> the MSF unit, the analyzed inefficiencies refer to fouling at different stages:<br />

• brine heater,<br />

• recovery section, <strong>and</strong><br />

• reject section<br />

Neither the MSF pumping process nor the brine level in each flash chamber were<br />

diagnosed since they were not simulated in the mathematical model. The <strong>analysis</strong><br />

could be performed with respect to thermal problems inside the distillers, vapor<br />

conditions to the brine heater or the TBT/distillate.<br />

As we will see in later sections, fouling in distillers was considered a global variable<br />

if it affected other distillers.<br />

The effect <strong>of</strong> these eight inefficiencies was measured on:<br />

• the behavior <strong>of</strong> the rest <strong>of</strong> the plant devices (intrinsic/induced malfunction <strong>and</strong><br />

dysfunction <strong>analysis</strong>),<br />

• additional fuel plant consumption (impact on fuel),<br />

• the thermoeconomic cost <strong>of</strong> electricity <strong>and</strong> distilled water,<br />

• the irreversibility increase <strong>of</strong> each unit.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> should cover as much <strong>of</strong> the maximum range <strong>of</strong> electricity<br />

<strong>and</strong> water production as possible so that intermediate dem<strong>and</strong>s can be predicted from<br />

204 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> diagnosis<br />

acquired experience. Four loads were considered under the most usual operating<br />

situations:<br />

• Full load in condensing mode (no extraction to MSF unit): 140 MW <strong>of</strong> power<br />

generated.<br />

• Full load in extraction mode (electricity <strong>and</strong> water production): 122 MW <strong>of</strong><br />

output power (89.68 kg/s <strong>of</strong> steam extracted to the desalination unit).<br />

• Partial load in extraction mode at 90 MW output power (60 kg/s steam extracted<br />

to the MSF unit).<br />

• Parallel mode (the reduction pressure station is opened to maintain the pressure<br />

to the MSF unit): 60 MW <strong>of</strong> output power (50 kg/s extraction to desalination).<br />

The first situation is a high-electricity dem<strong>and</strong> when the distiller has been stopped for<br />

repair, the two intermediate productions are the most common <strong>and</strong> the fourth is<br />

typical in winter.<br />

Two freshwater productions were analyzed under the following specific conditions:<br />

• 1,900 T/h distillate with 32 ºC seawater (the nominal production under Gulf<br />

seawater conditions in spring or autumn).<br />

• 2,400 T/h distillate with 25 ºC feedwater to the reject section (the maximum<br />

winter production). Seawater can be less than 25 ºC (the minimum temperature<br />

operation for the reject section), so the temper system uses a part <strong>of</strong> the reject<br />

cooling brine <strong>and</strong> stay secure in the last stage <strong>of</strong> the reject section.<br />

Loads <strong>and</strong> inefficiencies may be <strong>combined</strong> in many ways. We analyzed all <strong>of</strong> these<br />

possibilities but only present two: an inefficiency in the fourth section <strong>of</strong> the highpressure<br />

turbine <strong>and</strong> an inefficiency in the MSF unit (with the cleaning ball system in<br />

the heater) at a prefixed load. These examples represent a local <strong>and</strong> global variable in<br />

two separate systems. We subsequently considered the ‘upstream’ effect <strong>of</strong> fouling in<br />

the recovery section <strong>of</strong> the MSF plant on the steam power plant. Finally, the most<br />

general situation was analyzed when several inefficiencies in the power or<br />

desalination plant occurred together. The rest <strong>of</strong> the combinations (i.e. the <strong>analysis</strong> <strong>of</strong><br />

the individual inefficiencies presented above) are presented in Annex 1 for a 122 MW<br />

load in the power plant <strong>and</strong> the NTOS case <strong>of</strong> the MSF plant, including figures <strong>and</strong><br />

matrices calculated in the <strong>analysis</strong> <strong>of</strong> each inefficiency. The effect <strong>of</strong> the load in the<br />

above inefficiencies is summarized in section 7.3.4.<br />

7.3.2 Analysis <strong>of</strong> individual inefficiencies<br />

7.3.2.1 Inefficiency in the fourth section <strong>of</strong> the high-pressure turbine<br />

As defined by Royo (1994), an intrinsic malfunction in a steam turbine is expressed<br />

as the damage in the steam expansion process <strong>and</strong> energy transmission to the shaft<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 205


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

due to several factors including erosion, fractures, ruptures, sediments, surface finish,<br />

friction, steam path, seals <strong>and</strong> diaphragm deterioration, control valve <strong>and</strong> heat losses.<br />

An inefficiency can also be an induced malfunction due to the variation <strong>of</strong> external<br />

factors apart from component damage. These external factors include changes in<br />

admission temperature, exhaust pressure or extraction mass flows <strong>of</strong> a steam turbine<br />

(Zaleta, 1997).<br />

We simulated that the fourth section <strong>of</strong> the high-pressure turbine underwent behavior<br />

degradation <strong>and</strong>, as a result, the isoentropic efficiency decreased by 10% (at 122 MW<br />

total output power). The three upstream turbine sections are insensitive to <strong>and</strong><br />

incapable <strong>of</strong> responding to this inefficiency <strong>and</strong> the vapor conditions entering the<br />

inefficient section were maintained with respect to the design condition. A lower<br />

isoentropic efficiency means that the outlet steam vapor conditions have a higher<br />

enthalpy, if the exhaust pressure <strong>of</strong> the high-pressure turbine is controlled by the MSF<br />

system. Thus, the first induced malfunction in the MSF unit was due to the variation<br />

<strong>of</strong> external factors; the steam conditions entering the MSF plant were changed by an<br />

inefficiency (or intrinsic malfunction) in the fourth section <strong>of</strong> the high-pressure<br />

turbine.<br />

The output power <strong>of</strong> this section was also considerably lower because <strong>of</strong> the<br />

reduction in the enthalpy drop. The three sections <strong>of</strong> the high-pressure turbine<br />

maintained their power production. The steam pressure entering the low-pressure<br />

turbine was maintained <strong>and</strong> the exhaust pressure must be the same as in the design<br />

(we assumed that the ambient conditions remained unchanged <strong>and</strong> constant<br />

condenser pressure). Therefore, the efficiency <strong>of</strong> the low-pressure turbine should not<br />

vary considerably <strong>and</strong> the two sections <strong>of</strong> the low-pressure turbine do not produce<br />

additional power to maintain the final production.<br />

Consequently, additional live steam was needed to maintain final production. The<br />

three sections <strong>of</strong> the high-pressure turbine <strong>and</strong> the two sections <strong>of</strong> the low-pressure<br />

turbine provided the extra power not supplied by the inefficient section. The<br />

additional live steam affected the whole system, but the latter generally readapts to<br />

maintain design values: design feedwater system values were maintained by<br />

increasing the extraction mass flows. Pump consumption increased in proportion to<br />

the additional mass flow required by the boiler. As a result, no significant induced<br />

malfunctions were provoked by the inefficiency in the high-pressure turbine.<br />

The total impact on the fuel was 6.035 MW, but 6.015 MW in the inefficient<br />

component. Thus, the effect <strong>of</strong> the inefficiency could be considered local to the<br />

component with the intrinsic malfunction. Next we considered the contribution <strong>of</strong><br />

each component.<br />

The physical consequences <strong>of</strong> inefficiencies will be reviewed using the symbolic<br />

diagnosis notation <strong>of</strong> this Ph. D. Thesis (see Chapter 6 for nomenclature). The same<br />

methodology was used for each example. First the target conditions <strong>and</strong> the<br />

206 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> diagnosis<br />

inefficient situations were simulated (see Chapter 5). The design <strong>and</strong> inefficient<br />

situation include the most significant flowstreams, which are the basis <strong>of</strong> the<br />

thermoeconomic <strong>analysis</strong>. Following the F-P definitions adopted for the<br />

thermoeconomic model (section 7.1), the fuel <strong>and</strong> product table was prepared.<br />

table 7.23 corresponds to the design <strong>and</strong> table 7.24 to the inefficient condition. The<br />

unit exergy consumption κ <strong>of</strong> each component is very easy to calculate using the F-P<br />

tables (by dividing the fuels entering the plant by their product). Then the reference<br />

〈KP〉 matrix (table 7.25) <strong>and</strong> the 〈KP〉 matrix (table 7.26) are made for the inefficient<br />

mode. If these two matrices are subtracted, we obtain the ∆ 〈KP〉 matrix with the unit<br />

exergy consumption increase <strong>of</strong> each component (table 7.27). The ∆ 〈KP〉 matrix is<br />

the basis for calculating the endogenous irreversibility or malfunction. If the two<br />

matrices are multiplied, we obtain the irreversibility matrix |I〉 (table 7.28) with the<br />

irreversibility increase (or dysfunction coefficients) <strong>of</strong> each component. The first<br />

factor is the diagonal matrix K D–U D, where K D is the array containing the sum (by<br />

columns) <strong>of</strong> the 〈KP〉 matrix <strong>and</strong> U D is the unitary matrix. The second factor is the<br />

inverse <strong>of</strong> the unitary matrix minus the 〈KP〉 matrix, i.e., (U D–KP) –1 . The unit exergy<br />

cost <strong>of</strong> a product is the column sum <strong>of</strong> the dysfunction coefficients in the |I〉 matrix<br />

plus one (table 7.28). Finally, the dysfunction matrix [DF] needed to build the<br />

malfunction <strong>and</strong> dysfunction table is calculated by multiplying the |I〉 matrix by<br />

∆ 〈KP〉 P, where P is the array containing the product <strong>of</strong> each component. Thus, the<br />

irreversibility increase in each unit is connected to the increase in unit exergy<br />

consumption <strong>of</strong> each component. The malfunction <strong>of</strong> each component MF is the<br />

product ∆ 〈KP〉 P <strong>and</strong> is located at the end <strong>of</strong> the table. The column sum is the fuel<br />

impact <strong>of</strong> a component, i.e., the additional fuel plant consumption provoked by the<br />

considered unit <strong>and</strong> the row sum is the irreversibility increase <strong>of</strong> a component (see<br />

table 7.29).<br />

After having explained the most relevant matrices to analyze a plant inefficiency<br />

(table 7.29), we will now consider the results <strong>and</strong> explain the values using physical<br />

reasons. Figure 7.13 shows the impact on fuel <strong>analysis</strong> from the malfunction/<br />

dysfunction table (included in table 7.29) <strong>and</strong> figure 7.14 includes the irreversibility<br />

increase <strong>of</strong> each component <strong>of</strong> the power plant.<br />

The intrinsic malfunction is the easiest to explain. When the fourth section <strong>of</strong> the<br />

high-pressure turbine was working at 10% less isoentropic efficiency than normal, the<br />

output power (P in the F-P table 7.24) decreased but the section's steam conditions<br />

were maintained. The irreversibility increased (the turbine section increased its<br />

irreversibility to 3,270 kW, table 7.29), <strong>and</strong> the resources required to produce the<br />

same output power increased as well as the unit exergy consumption <strong>of</strong> the<br />

component ∆k (∆k = 0.2144, see table 7.27). Multiplying by the product in this<br />

section (19.23 MW), the malfunction was 4.12 MW (see table 7.29). The fuel impact<br />

due to the inefficient component was 6.01 MW (see also the table 7.29).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 207


FIGURE 7.13 Impact on fuel <strong>analysis</strong> when the efficiency <strong>of</strong> the HPT4 is decreased 10%.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

208 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

FIGURE 7.14 Irreversibility increase <strong>analysis</strong> with the inefficiency in the HPT4.


TABLE 7.23 F-P diagram in design, output power <strong>of</strong> 122 MW .<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 209


TABLE 7.24 F-P values with inefficiency in HPT4 (10% lower efficiency).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

210 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.25 KP matrix in design (122 MW).<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 211


TABLE 7.26 KP matrix with inefficiency in HPT4 (10%).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

212 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.27 Variation de KP with inefficiency in HPT4.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 213


TABLE 7.28 Irreversibility matrix I with an inefficiency in HPT4.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

214 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.29 Dysfunction/malfunction matrix with inefficiency in HPT4 (10% isoentropic eff.).<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 215


TABLE 7.30 Malfunction matrix with inefficiency in HPT4 (1% isoentropic eff. is varied).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

216 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> diagnosis<br />

As mentioned, the inefficiency also affected MSF unit behavior. This induced<br />

malfunction was expected because the steam leaving the HPT4 section is consumed<br />

in the MSF unit. Since the MSF product (exergy flow <strong>of</strong> distilled water) is constant,<br />

the variation <strong>of</strong> the steam conditions entering the MSF unit directly affects its<br />

behavior (we assumed that the condensate returned to the deaerator maintains its<br />

properties independent <strong>of</strong> inlet conditions). A higher enthalpy in the exhaust vapor <strong>of</strong><br />

the high-pressure turbine should imply a higher specific consumption per freshwater<br />

unit produced, as seen in the variation <strong>of</strong> the unit exergy consumption (∆k = 0.075,<br />

see the corresponding value in table 7.27). But the thermoeconomic model gives an<br />

important function to the MSF unit: the negentropy generated in the MSF heater. The<br />

inefficiency in the fourth section <strong>of</strong> the high-pressure turbine generated a higher<br />

negentropy in the MSF unit (the entropy <strong>of</strong> exhaust vapor from the turbine increases<br />

with a lower isoentropic efficiency). This negentropy is a secondary product <strong>of</strong> the<br />

MSF unit. Its increase implies a decrease in unit exergy consumption <strong>of</strong> the<br />

component (the ∆k variation due to negentropy generation is –0.154, see table 7.27).<br />

Balancing the two terms, the increase in unit exergy consumption in the MSF was<br />

negative, provoking –537 kW induced malfunction. In conclusion, the value <strong>of</strong> the<br />

induced malfunction in this component was due to the thermoeconomic model. It did<br />

not correspond to the expected response to an intrinsic malfunction in the fourth<br />

section <strong>of</strong> the high-pressure turbine. In other words, the negentropy generated in the<br />

MSF unit reduced the cost <strong>of</strong> water because the negentropy generated in the MSF unit<br />

reduced the cost <strong>of</strong> the condenser.<br />

The physical <strong>analysis</strong> <strong>of</strong> the inefficiency did not detect any more induced<br />

malfunctions in the system, although two components had a higher induced<br />

malfunction than the accuracy <strong>of</strong> the simulator: the boiler (–128 kW) <strong>and</strong> the first<br />

section <strong>of</strong> the HPT (–331 kW). These values are the consequence <strong>of</strong> a very high<br />

component product since unit exergy consumption increase was almost zero in both<br />

cases. This consumption varied only slightly because the steam needed to produce the<br />

required power increased with the simulated inefficiency.<br />

The irreversibility increase in each component (table 7.29) was calculated by<br />

subtracting the fuel-product differences in tables 7.23 <strong>and</strong> 7.24, or by adding the unit<br />

malfunction to the unit dysfunction generated by the malfunction <strong>of</strong> the rest <strong>of</strong> units<br />

in the system. In our example, the boiler dysfunction was the highest, mainly due to<br />

the malfunctions in HPT1, HPT4 <strong>and</strong> MSF (see table 7.29), the most important ones<br />

detected in this case. The dysfunction generated in the condenser was also important,<br />

but the cause was again the three components undergoing the malfunction. Boiler <strong>and</strong><br />

condenser production increased by about 3 MW (this additional production was<br />

required by the rest <strong>of</strong> components to maintain the final production <strong>of</strong> the steam<br />

power plant with the inefficiency simulated in the fourth section <strong>of</strong> the HPT). In the<br />

productive structure (figure 7.5), the two products generated by these two<br />

components (the availability <strong>of</strong> the steam generated in the boiler <strong>and</strong> the negentropy<br />

generated in the steam cycle) were easily apportioned to the rest <strong>of</strong> the plant<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 217


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

components. Figure 7.13 shows the irreversibility increase <strong>analysis</strong> <strong>of</strong> this<br />

inefficiency.<br />

Some explanations are required regarding the malfunction <strong>and</strong> dysfunction values <strong>of</strong><br />

the non-physical components <strong>of</strong> our thermoeconomic model. A junction is a nonphysical<br />

device <strong>and</strong> is fictitious in the productive structure. Its function, similar to that<br />

<strong>of</strong> branching points, is structural, i.e. junctions <strong>and</strong> branches show how the resources<br />

are distributed among the plant devices. The malfunction <strong>and</strong> the dysfunction<br />

generated in a junction must be zero: in equation (6.45) the unit exergy consumption<br />

increase in a junction is zero <strong>and</strong> the dysfunction coefficients φ responsible for the<br />

dysfunction generated by other components are also zero (remember that the<br />

dysfunction coefficients φ only depend on the unit exergy consumption k <strong>of</strong> the<br />

component in operating conditions). However, a junction can generate a dysfunction<br />

in other system components (see equation 6.46). The value <strong>of</strong> the dysfunction<br />

strongly depends on the dysfunction coefficients φ <strong>of</strong> each component where the<br />

dysfunction is generated. For example, the unit exergy consumption k <strong>of</strong> junction J4<br />

varies with a change in the unit exergy consumption <strong>of</strong> its exergy ratios r (the<br />

electricity produced in the turbine sections). All the boiler φ coefficients were nonnegative<br />

(the dysfunction generated by the junction in the boiler was not zero, –445<br />

kW). The junction usually generates dysfunctions due to the variation in the fuels<br />

(i.e., the product <strong>of</strong> the units that enter the junction) but the components that have<br />

non-zero values in all their φ coefficients also suffer from the junction dysfunction.<br />

These special components are the boiler <strong>and</strong> condenser, which are interrelated with<br />

the rest <strong>of</strong> components in the productive structure <strong>of</strong> the power plant (see figure 7.5).<br />

The impact on fuel <strong>analysis</strong> is similar to the previous <strong>analysis</strong>, but here the impact on<br />

fuel consumption is the sum <strong>of</strong> the malfunction <strong>and</strong> the dysfunction generated by<br />

each component in all others (see figure 7.14). Logically, the dysfunctions generated<br />

by HPT1, HPT4 <strong>and</strong> MSF in the boiler <strong>and</strong> the condenser were the most important.<br />

One <strong>of</strong> the most useful applications <strong>of</strong> the thermoeconomic diagnosis is the<br />

malfunction matrix. It provides information about the malfunction associated with<br />

each component during an inefficiency. It is a very valuable tool to predict system<br />

behavior without using the simulator (recall that the same results were obtained using<br />

either the diagnosis or simulator). We want to predict the additional fuel consumption<br />

with an inefficiency <strong>and</strong> maintain the equations that model the physical behavior <strong>of</strong><br />

the plant in the simulator (performing each individual <strong>analysis</strong> for an operating<br />

condition). At least two premises are required to create the malfunction matrix:<br />

The response <strong>of</strong> the system must be proportional to the degree <strong>of</strong> inefficiency (impact<br />

on fuel, associated malfunctions, etc.). To calculate the fuel impact <strong>of</strong> a known<br />

inefficiency, the corresponding malfunction matrix need only be multiplied or<br />

divided, depending on the ratio <strong>of</strong> the real inefficiency <strong>and</strong> the inefficiency defined in<br />

the malfunction matrix.<br />

218 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> diagnosis<br />

To predict the effect <strong>of</strong> several malfunctions, the inefficient components must be local<br />

to their subsystems. The total impact on fuel can then be calculated as the sum <strong>of</strong> the<br />

malfunction matrices associated with the individual inefficiencies.<br />

The second assumption is not necessary here because we only analyzed an individual<br />

inefficiency. The first premise could be checked by analyzing the graphic impact on<br />

fuel <strong>analysis</strong> versus the degree <strong>of</strong> inefficiency applied. In this case, the isoentropic<br />

efficiency <strong>of</strong> the fourth section <strong>of</strong> HPT was varied from –10% to +10% with respect<br />

to the design efficiency (around 85%). Figure 7.15 shows how the linearity <strong>of</strong> the<br />

sensitivity <strong>analysis</strong> varies while the plant load is kept constant (122 MW <strong>of</strong> output<br />

power in extraction mode <strong>and</strong> 2,400 T/h freshwater production).<br />

FIGURE 7.15 Additional fuel consumption when varying the isoentropic efficiency in HPT4.<br />

Inc. fuel consumption<br />

6000<br />

kW<br />

4000<br />

2000<br />

-10 -8 -6 -4 -2<br />

-2000<br />

0 2 4 6 8 10<br />

-4000<br />

-6000<br />

0<br />

% eff. in HT4<br />

Plant behavior was linear when we varied this inefficiency (figure 7.15). The<br />

malfunction matrix in table 7.30 is very useful to calculate the malfunctions<br />

associated with each inefficiency (by summing the columns <strong>and</strong> multiplying each<br />

component by its product). The high unit exergy consumption <strong>of</strong> the condenser pump<br />

was the result <strong>of</strong> the mathematical model (as were the high values <strong>of</strong> the low-pressure<br />

heater no. 2). In these two cases, the low product values minimized the previously<br />

mentioned effect in the malfunction <strong>analysis</strong>. The MSF components <strong>of</strong> this matrix<br />

were very high but the low exergy value <strong>of</strong> its product (freshwater) induced a low<br />

malfunction. All sections <strong>of</strong> the high-pressure turbine were affected by the<br />

inefficiency but, as expected, the fourth section had the highest value. The values <strong>of</strong><br />

the first section <strong>of</strong> the high <strong>and</strong> low-pressure turbine were also considerable since<br />

they had to readapt their products to maintain final production.<br />

The effect <strong>of</strong> the inefficiency can be quantified as the total cost (including capital cost<br />

<strong>of</strong> devices) <strong>of</strong> electricity <strong>and</strong> water, which is especially illustrative for plant<br />

managers. Electricity increases 0.000033 $/kWh per 1% variation in efficiency<br />

(figure 7.16) or a yearly savings <strong>of</strong> 35,200 $/y.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 219


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

FIGURE 7.16 Unit electricity cost when the isoentropic HPT4 efficiency is modified.<br />

Electricity cost<br />

0,0383<br />

$/kWh<br />

0,0381<br />

0,0379<br />

0,0377<br />

0,0375<br />

0,0373<br />

% eff. in HT4<br />

-10 -8 -6 -4 -2 0 2 4 6 8 10<br />

Surprisingly, the effect on the cost <strong>of</strong> water was even greater –in absolute terms- than<br />

for electricity (0.00047 $/m 3 per 1% inefficiency, or almost 10,000 $/y; figure 7.17),<br />

although the relative cost <strong>of</strong> electricity varied more. This is because the apparently<br />

local inefficiency changes the steam conditions sent to MSF unit, which implies an<br />

additional cost, mainly due to the high exergetic cost associated with water (see<br />

table 7.28).<br />

FIGURE 7.17 Unit distilled water cost when the isoentropic HPT4 efficiency is modified.<br />

Water cost<br />

1,278<br />

$/m3<br />

1,274<br />

1,270<br />

1,266<br />

% eff. in HT4<br />

-10 -8 -6 -4 -2 0 2 4 6 8 10<br />

The main conclusions <strong>of</strong> our <strong>analysis</strong> <strong>of</strong> an inefficiency in the final section <strong>of</strong> the<br />

high-pressure turbine are:<br />

• The isoentropic efficiency only affected the behavior <strong>of</strong> the inefficient<br />

component <strong>and</strong> provoked a small malfunction in the MSF plant by changing<br />

exhaust vapor conditions leaving the HPT.<br />

220 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> diagnosis<br />

• The steam power plant could not readapt its behavior to maintain the final<br />

production. Additional live steam was required to produce the electricity<br />

dem<strong>and</strong>ed, consuming more fuel (6,035 kW). The dysfunction <strong>analysis</strong> was<br />

useful to observe how the components that provide the energy quality to the<br />

steam cycle (boiler <strong>and</strong> condenser) have to increase their productions that are<br />

distributed to the rest <strong>of</strong> plant components, producing the additional power not<br />

supplied by the inefficient section <strong>of</strong> the turbine.<br />

• The effect <strong>of</strong> this inefficiency was quite significant <strong>and</strong> represented an additional<br />

water <strong>and</strong> electricity cost <strong>of</strong> 0.00047 $/m 3 <strong>and</strong> 0.000033 $/kWh respectively, per<br />

unit <strong>of</strong> efficiency (or 45,200 $/y in both products). The nature <strong>of</strong> the inefficiency<br />

should be studied carefully, taking into account several factors including repair<br />

time, personnel costs <strong>and</strong> the price <strong>of</strong> the components if they need to be replaced<br />

to avoid extra natural gas consumption.<br />

• The sensitivity <strong>analysis</strong> applied in a reasonable range revealed a linear response<br />

by the simulator mathematical model. Thus, the malfunction matrix can substitute<br />

new <strong>simulation</strong>s with this inefficiency <strong>and</strong> predict its effect on a real plant.<br />

• The value <strong>of</strong> the induced MSF unit malfunction demonstrates that plant<br />

diagnosis strongly depends on the thermoeconomic model. Sometimes the<br />

physical consequences <strong>of</strong> an inefficiency cannot be translated into a table <strong>of</strong><br />

expected values for fuel impact or irreversibility increase <strong>of</strong> a process or<br />

component.<br />

7.3.2.2 Using the cleaning ball system in the brine heater<br />

The fouling resistance R f (for definition see section 3.2.1) involves three resistances:<br />

• Resistance due to fouling or scale inside the tube.<br />

• Resistance due to fouling outside the tube.<br />

• Resistance due to the accumulation <strong>of</strong> non-condensable gases in the vapor.<br />

The cleaning ball system can only reduce tube fouling or scale in a heat exchanger,<br />

one <strong>of</strong> the main causes <strong>of</strong> performance loss in MSF plants in the high-temperature<br />

sections. In general, fouling occurs when deposits are laid down on the heat transfer<br />

surfaces (Hanbury, Hodgkiess <strong>and</strong> Morris, 1993). These deposits can be due to scale<br />

from the reverse solubility <strong>of</strong> salts in the brine, dirt from corrosion products or<br />

biological growths on the surface. The latter only occurs in the rejection section <strong>and</strong><br />

can be controlled by feed chlorination. The scale type depends on the brine chemistry,<br />

plant conditions, chemical additives to the feed <strong>and</strong> the type <strong>of</strong> cleaning. In general,<br />

calcium carbonate <strong>and</strong> calcium sulfate are the most common forms <strong>of</strong> scale.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 221


TABLE 7.31 F-P values (design) for the MSF plant. Nominal production in summer.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

222 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.32 F-P values without fouling in heater. Nominal production, 32 ºC seawater.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 223


TABLE 7.33 KP matrix in design.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

224 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.34 KP matrix without fouling in heater. NTOS data case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 225


TABLE 7.35 Variation <strong>of</strong> the KP matrix without fouling in heater. NTOS case.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

226 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.36 Irreversibility matrix without fouling in heater. 1,900 T/h <strong>and</strong> 32 ºC seawater temp.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 227


TABLE 7.37 Malfunction/dysfunction matrix without fouling in heater. NTOS case.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

228 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.38 Malfunction matrix varying fouling in heater 0,00001 m2 K/W in NTOS case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 229


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

The fouling effect in the MSF plant was simulated, quantified <strong>and</strong> analyzed for the<br />

brine heater. The cleaning ball system was assumed to be working at maximum <strong>and</strong><br />

fouling in brine heater was set to zero (fouling factor in heater at design conditions<br />

was 0.00025 m 2 ·K/W), although this is impossible in practice since outside fouling<br />

<strong>and</strong> non-condensable gas phenomena cannot be avoided. The reference case had the<br />

same operating conditions without the cleaning ball system. For this reason, most<br />

malfunctions associated with the cleaning ball system are negative (they should be<br />

called ‘benefunctions’), i.e., they save fuel. The malfunction <strong>analysis</strong> was performed<br />

at 1,900 T/h water production with 32 ºC seawater (the first <strong>of</strong> the two examples).<br />

Water production was constant although but this does not imply a constant product<br />

exergy flow.<br />

To explain how fouling in the brine heater affects MSF behavior, first the recycle<br />

brine, seawater to reject <strong>and</strong> make-up flows (R, SR, F) were maintained at designed<br />

levels. The condensation temperature <strong>of</strong> the steam provided by the steam power plant<br />

also remained constant. A lower fouling inside the brine heater improved the overall<br />

heat transfer coefficient, which implied that:<br />

• The interstage temperature difference in the heater was reduced, i.e. the cooling<br />

brine temperatures entering (TF,1) <strong>and</strong> leaving (TBT = TB,O) the heater were<br />

increased.<br />

• The temperature rise <strong>of</strong> the cooling brine in the heater was also increased.<br />

A higher Top Brine Temperature (TBT) implies a higher flash range ∆T <strong>and</strong> more<br />

freshwater production. The temperature pr<strong>of</strong>ile <strong>of</strong> the recovery <strong>and</strong> reject section is<br />

altered if the temperatures entering <strong>and</strong> leaving the recovery section are increased. If<br />

the final production is to be maintained, R, SR <strong>and</strong> F must be decreased. But even the<br />

TBT <strong>and</strong> T F,1 reach higher than design temperatures (<strong>and</strong> therefore the temperatures<br />

pr<strong>of</strong>ile in recovery <strong>and</strong> reject sections). Brine fouling is a global variable in the MSF<br />

unit since it affects the rest <strong>of</strong> the system.<br />

About 1,411 kW <strong>of</strong> fuel was saved with the benefunction in different plant<br />

components (not only in the heater). Less steam was consumed, affecting the<br />

behavior <strong>of</strong> the steam power plant when less steam is required for this extraction, as<br />

in the next example.<br />

Inefficiency was diagnosed using the symbolic notation explained in Chapter 6. The<br />

simulator in Chapter 5 was used to obtain the F <strong>and</strong> P values for the reference<br />

conditions <strong>and</strong> inefficient situation. Following the productive structure <strong>of</strong> the MSF<br />

unit (see figure 7.11) with 1,900 T/h water production, the F <strong>and</strong> P values are<br />

included in tables 7.31 <strong>and</strong> 7.32 respectively, using the nomenclature in table 7.4 for<br />

the components. In this case, the matrix was 18×18 (11 components <strong>and</strong> 7 junctions)<br />

whereas the matrix was 30×30 (26 components <strong>and</strong> 4 junctions) in the power plant<br />

<strong>analysis</strong>. The ∆ 〈KP〉 matrix (table 7.35) was built by subtracting the 〈KP〉 reference<br />

matrix (table 7.33) <strong>and</strong> the 〈KP〉 matrix (table 7.34) corresponding to an inefficient<br />

230 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> diagnosis<br />

operation. The latter were obtained by dividing the F-P tables. The irreversibility<br />

matrix |I〉 (table 7.36) contains the irreversibility <strong>and</strong> unit exergy costs <strong>of</strong> each<br />

component. Finally, the dysfunction table (table 7.37) contains the dysfunction<br />

coefficients φ <strong>and</strong> the malfunction array MF. The column sum is the fuel impact <strong>of</strong> a<br />

component <strong>and</strong> the row sum is the irreversibility increase <strong>of</strong> a component. Figure<br />

7.18 shows the impact on fuel <strong>analysis</strong> in the malfunction/dysfunction table. Figure<br />

7.19 includes the irreversibility increase <strong>of</strong> each component <strong>of</strong> the power plant.<br />

Having obtained the dysfunction matrix (which provides information about the state<br />

<strong>of</strong> a given plant with an inefficiency), we analyzed the malfunctions in the<br />

desalination plant components. The physical variations in the MSF plant with the<br />

benefunction were translated into malfunctions. The first important conclusion is that<br />

the malfunction generated in the brine heater was not the highest. The malfunction<br />

induced in other components was more important than the intrinsic malfunction<br />

provoked by heater inefficiency. Therefore, each malfunction should be analyzed<br />

separately.<br />

The intrinsic malfunction is quite easy to explain. Using the cleaning ball system in<br />

the heater improves the heat transfer process in the tubes. This reduces the thermal<br />

irreversibility, assuming that the mechanical <strong>and</strong> chemical irreversibility is<br />

maintained. The irreversibility was reduced by 865 kW in this component (see table<br />

7.37), increasing its exergetic efficiency. The reference unit exergy consumption was<br />

reduced with respect to the inefficient condition (respectively 1.096 <strong>and</strong> 1.075 in<br />

tables 7.33 or 7.34), or the change in unit exergy consumption ∆k decreased with<br />

respect to the reference state. The decrease <strong>of</strong> the unit exergy consumption (–0.02) is<br />

included in the ∆ 〈KP〉 matrix (table 7.35). The product <strong>of</strong> the heater is the cooling<br />

brine heated to the TBT (42,021 kW), then the intrinsic malfunction <strong>of</strong> –875 kW. The<br />

impact on fuel saved in this component was 2,419.6 kW (both values are in<br />

table 7.37).<br />

The induced malfunction in the recovery section was positive (203 kW) <strong>and</strong> the<br />

irreversibility increase was 247 kW in the process (see both values in table 7.37).<br />

Consequently, the variation in unit energy consumption in this component with heater<br />

fouling was positive (∆k = 0.025, see table 7.35), i.e. the process was more inefficient<br />

in this section. Assuming that the distillate quantity <strong>and</strong> quality is maintained, an<br />

uncontrolled TBT increases due to the effect <strong>of</strong> the cleaning ball system in the brine<br />

heater. Although cooling brine was also increased, the temperature rise was lower<br />

than the TBT (because <strong>of</strong> the two effects <strong>of</strong> fouling in the brine heater). Thus, the<br />

amount <strong>of</strong> energy needed to produce the distillate in the recovery section was higher<br />

than in the design situation. The efficiency <strong>of</strong> the component decreased <strong>and</strong> provoked<br />

an additional fuel consumption <strong>of</strong> 494 kW.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 231


FIGURE 7.18 Impact on fuel <strong>analysis</strong> when the fouling in BH is neglected.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

232 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

FIGURE 7.19 Irreversibility increase in the MSF with BH=0. NTOS case.


<strong>Thermoeconomic</strong> diagnosis<br />

On the other h<strong>and</strong>, the temperature pr<strong>of</strong>ile in the reject section remained almost<br />

unchanged because the effect <strong>of</strong> the heater fouling is far away from the reject section.<br />

A higher TBT also implies lower recycled brine R flowing toward the reject section to<br />

maintain final production. This flow is the main contribution <strong>of</strong> the reject section to<br />

produce distilled water, which was maintained constant (the reject section product is<br />

practically the final product <strong>of</strong> the MSF unit). Less energy was needed to produce<br />

freshwater. The efficiency was increased <strong>and</strong> the variation <strong>of</strong> the unit exergy<br />

consumption <strong>and</strong> irreversibility generated were reduced in the inefficient case (∆k = –<br />

0.013, ∆I = –91 kW, resulting in a negative malfunction <strong>of</strong> 91 kW (see tables 7.35 <strong>and</strong><br />

7.37) <strong>and</strong> 725 kW in fuel savings.<br />

The induced malfunction associated with the mixer was quite substantial (-- 942 kW).<br />

The make-up F <strong>and</strong> recirculation R flows were decreased by the cleaning ball system<br />

in the heater under constant final production <strong>of</strong> freshwater. The mechanical <strong>and</strong><br />

thermal irreversibility <strong>of</strong> the mixing process is logically reduced if the two flows<br />

entering the mixing chamber are reduced. The unit exergy consumption <strong>of</strong> the<br />

process or the irreversibility increase was reduced (∆k = - 0.0159, ∆I = –912 kW, see<br />

tables 7.35 <strong>and</strong> 7.37) <strong>and</strong> 1,360 kW (table 7.37) <strong>of</strong> fuel was saved.<br />

The fictitious device is a non-physical component intercalated at the beginning <strong>of</strong> the<br />

productive structure <strong>of</strong> the MSF unit (see figure 7.11). It charges the exergy costs <strong>of</strong><br />

the distiller flows with the plant residues: brine blowdown <strong>and</strong> reject cooling<br />

seawater. There is no physical explanation for malfunction <strong>of</strong> this device but the<br />

thermoeconomic model suggests two causes:<br />

• The exergy flow <strong>of</strong> the residues is higher (the fuel <strong>of</strong> this unit). The specific<br />

energy or mass flow rate <strong>of</strong> one <strong>of</strong> the two streams must be increased by an<br />

inefficiency in the MSF unit.<br />

• The steam to the brine heater decreases (here the unit product corresponds to the<br />

fuel <strong>of</strong> the brine heater).<br />

The second cause provoked a positive malfunction <strong>of</strong> 938 kW in the FD <strong>and</strong><br />

1,222 kW <strong>of</strong> extra fuel consumption (table 7.37). The same MSF residues are sent out<br />

to sea at a higher cost to the distiller when less fuel is consumed to produce water.<br />

The amount <strong>of</strong> irreversibility in each component is the sum <strong>of</strong> its own malfunction<br />

plus the dysfunctions generated by the malfunction <strong>of</strong> other components. Only the<br />

fictitious device had a considerable dysfunction value (–764 kW), generated by<br />

malfunctions in the brine heater, recovery <strong>and</strong> reject sections, mixer <strong>and</strong> several<br />

junctions (see table 7.37). As above (tables 7.31 <strong>and</strong> 7.32), the product <strong>of</strong> the<br />

fictitious device decreased more than 700 kW to readapt the use <strong>of</strong> the cleaning ball<br />

system in the brine heater under constant freshwater production. The dysfunction<br />

depends on the φ coefficients <strong>of</strong> the component. Since the fictitious device is at the<br />

beginning <strong>of</strong> the productive structure, most <strong>of</strong> its φ coefficients were non-zero values.<br />

In conclusion, dysfunction <strong>analysis</strong> is clearly unrelated to the physical behavior <strong>of</strong> the<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 233


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

plant, i.e., the MSF components do not vary production to maintain the final distillate<br />

due to the malfunctions.<br />

The impact on fuel <strong>analysis</strong> was similar to the previous <strong>analysis</strong>. In this case, the<br />

impact on fuel consumption was the sum <strong>of</strong> the malfunction <strong>and</strong> dysfunction<br />

generated by each component on others (table 7.37). As expected, the dysfunction<br />

generated by the brine heater, recovery <strong>and</strong> reject section <strong>and</strong> the mixer are important.<br />

Note that the malfunction/dysfunction <strong>analysis</strong> considers an unchanged final product.<br />

This is quite easy when the product is electricity, since the simulator can control the<br />

power output. However, the exergy flow <strong>of</strong> freshwater as the final product has two<br />

terms: the mass flow <strong>and</strong> the specific exergy <strong>of</strong> water leaving the distiller unit<br />

(quantity * quality, see Structural Theory, Valero et al., 1993). The mass flow must be<br />

controlled in the simulator but the specific distillate exergy is a function <strong>of</strong> the<br />

distiller temperature. The latter temperature depends on the operating conditions <strong>of</strong><br />

the MSF unit: seawater temperature <strong>and</strong> concentration, fouling in each section, etc. In<br />

our example, the water temperature leaving the distillate pump did not vary with the<br />

brine heater fouling. If the temperature changed, the impact on fuel associated with<br />

the change in total production is k * ∆P, where k * is the exergy cost <strong>of</strong> the product <strong>and</strong><br />

∆P is the variation <strong>of</strong> the total production (the value is shown in the right-bottom<br />

corner <strong>of</strong> the DF/MF table). The impact on fuel associated with the variation <strong>of</strong> the<br />

final product can be more important than the impact on fuel associated with the<br />

variation <strong>of</strong> the unit exergy consumption ∆k in each component (the total contribution<br />

due to both variations is also shown at the end <strong>of</strong> the DF/MF table).<br />

Having explained the most important results <strong>of</strong> MSF plant diagnosis without heater<br />

fouling, we can consider one <strong>of</strong> the most useful applications. Figure 7.20 can be used<br />

to study the linearity <strong>of</strong> the simulator (<strong>and</strong> a real plant, since the simulator was<br />

validated using data collected from a physical plant) to validate the malfunction<br />

matrix. Changing the design fouling factor in the brine heater (25×10 –5 m 2 ·K/W)<br />

gradually to zero saves fuel when the plant was operating to produce the same<br />

quantity <strong>of</strong> water as in the example.<br />

The model was reasonably linear when heater fouling was varied, at least for nominal<br />

production conditions in summer. However, at maximum operation, some internal<br />

flows like the recirculation flow R, make-up F or seawater to reject SR reached a<br />

maximum <strong>and</strong> the effect <strong>of</strong> the cleaning ball system was lower than expected for that<br />

load.<br />

Table 7.38 shows the malfunction matrix associated with each component when the<br />

fouling factor in the brine heater was changed by 0.00001 m 2 K/W. The most<br />

important terms <strong>of</strong> the matrix are associated with the above mentioned components:<br />

fictitious device, brine heater, recovery <strong>and</strong> reject sections <strong>and</strong> the mixer. These values<br />

can also be explained by analyzing the malfunctions associated with this inefficiency.<br />

As expected, pumps were not affected by brine heater fouling. The impact on fuel due<br />

234 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> diagnosis<br />

to changes in brine heater fouling can be calculated by multiplying the components <strong>of</strong><br />

the malfunction matrix by the product <strong>of</strong> each component <strong>and</strong> their unit exergy cost<br />

(obtained from the irreversibility matrix). Sometimes the malfunction matrix has<br />

components with high values, but the low product or low exergy cost associated with<br />

this component results in a lower impact on fuel.<br />

FIGURE 7.20 Impact on fuel <strong>analysis</strong> when the fouling in heater is varied.<br />

-500<br />

-1000<br />

-1500<br />

-2000<br />

-2500<br />

-3000<br />

Inc. fuel consumption<br />

Knowing the monetary cost <strong>of</strong> fresh water as a function <strong>of</strong> an inefficiency helps plant<br />

managers take decisions on using the cleaning ball system, depending on the<br />

compromise between consumption, operating costs <strong>and</strong> energy saved. Note that the<br />

cost <strong>of</strong> water decreased when heater fouling was decreased (figure 7.21).<br />

FIGURE 7.21 Monetary cost <strong>of</strong> distillate when the fouling in heater is varied.<br />

1,475<br />

1,470<br />

1,465<br />

1,460<br />

0<br />

fouling*10-5 in BH<br />

0 5 10 15 20 25<br />

kW<br />

$/m3<br />

Water cost<br />

fouling*10-5 in BH<br />

0 5 10 15 20 25<br />

In the nominal case, 0.00045 $/m 3 was saved when fouling was decreased by<br />

10 -- 5 m 2 ·K/W (or 7,650 $/y). Although the effect <strong>of</strong> the cleaning ball system was very<br />

difficult to translate into a constant fouling variation, the system reduced the fouling<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 235


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

factor several times over (see section 3.6.1) when the cleaning system was<br />

periodically connected (for example in a four-hour cycle). At maximum production,<br />

the cost decreases due to the effect <strong>of</strong> purchase costs <strong>and</strong> because the increase in<br />

exergy cost is lower than in the nominal case (remember that the internal flows reach<br />

a maximum during maximum winter production).<br />

The sensitivity <strong>analysis</strong> <strong>of</strong> the monetary cost <strong>and</strong> fuel impact takes into account that<br />

the exergy costs k * <strong>of</strong> the steam to the brine heater <strong>and</strong> the electricity for the MSF<br />

pumps are different from unity (unit exergy cost <strong>of</strong> steam to heater <strong>and</strong> vacuum system<br />

was 2.55 <strong>and</strong> 2.5 respectively <strong>and</strong> exergy cost <strong>of</strong> electricity was 2.85). So, the real cost<br />

<strong>of</strong> producing water <strong>and</strong> the consequences <strong>of</strong> an inefficiency can be dealt with correctly.<br />

In summary, using the cleaning ball system in the heater had the following<br />

consequences:<br />

• It changed the temperature pr<strong>of</strong>iles <strong>of</strong> the cooling, flashing brine <strong>and</strong> distillate in<br />

the MSF unit. As the brine heater is settled at the beginning <strong>of</strong> the process, the<br />

temperatures <strong>of</strong> the recycle brine before <strong>and</strong> after the heater were affected. As<br />

those temperatures enter <strong>and</strong> leave the recovery section, the whole system was<br />

influenced by this inefficiency.<br />

• As a consequence <strong>of</strong> the last point, the induced malfunctions in the rest <strong>of</strong><br />

components were higher than the intrinsic malfunction in the heater, taking into<br />

account the dimensions <strong>of</strong> each component. However, the dysfunction <strong>analysis</strong><br />

did not provide any interesting information on how the components readapted<br />

their production to maintain the final production <strong>of</strong> freshwater. The non-physical<br />

components cannot be explained from a physical viewpoint.<br />

• The model was linear under changes in heater fouling. So, the malfunction<br />

matrix can be used to predict the fuel saved with the cleaning ball system or<br />

component malfunctions.<br />

• The cleaning ball system in heater increased the TBT <strong>of</strong> the unit. This implies a<br />

lower consumption to produce the same amount <strong>of</strong> freshwater, but also provokes<br />

scale formation due to the high-operation temperatures. Consequently, the<br />

cleaning ball system should be continuously maintained in the heater to keep the<br />

fouling factor low. If the system is not operating, scale formation will reduce the<br />

effectiveness <strong>of</strong> the condenser <strong>and</strong> the whole MSF unit.<br />

7.3.2.3 The effect <strong>of</strong> recovery section fouling on steam power plant behavior<br />

An inefficiency in a power plant or desalination unit will provoke additional fuel<br />

consumption. The <strong>analysis</strong> was performed separately for both plants. But if an<br />

inefficiency in the MSF unit provokes an increase/decrease in steam consumption by<br />

the brine heater, how does the steam power plant readapt?. If the electricity <strong>and</strong> water<br />

production are kept constant, the inefficiency in the MSF unit is an inlet parameter<br />

that seriously affects power plant behavior. This parameter is the amount <strong>of</strong> steam<br />

236 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> diagnosis<br />

diverted to the MSF unit. Our <strong>analysis</strong> considered an inefficiency detected<br />

‘downstream’ <strong>and</strong> the MSF unit can induce the malfunctions; an inefficiency detected<br />

downstream should also be quantified upstream. So, we considered the ‘co-lateral’<br />

effects <strong>of</strong> a co-generation plant with this example.<br />

In this example, electricity production was held at 122 MW in MCR operating<br />

conditions. The fouling in the recovery section was reduced to zero by the cleaning<br />

ball system <strong>and</strong> the live steam leaving the boiler was maintained. The first physical<br />

consequence (the effects <strong>of</strong> the cleaning ball system in the recovery section are<br />

explained in section 6 <strong>of</strong> Annex 1) <strong>of</strong> an inefficiency was a reduction in steam<br />

consumption from 89.1 to 71.1 kg/s (corresponding to a freshwater production <strong>of</strong><br />

2,400 T/h, the maximum distillated in a MSF unit). Extraction to the MSF unit is at<br />

the end <strong>of</strong> the high-pressure turbine, so the latter was not affected by the different<br />

uses <strong>of</strong> the exhaust steam from this turbine. If the steam leaving the high-pressure<br />

turbine is not diverted to the MSF unit when some <strong>of</strong> it is saved with the cleaning ball<br />

system, an extra quantity <strong>of</strong> steam goes to the low-pressure turbine. Although the<br />

final section <strong>of</strong> the turbine has to maintain the exhaust pressure (we maintain the<br />

external parameters <strong>of</strong> the plant), at least the efficiency <strong>of</strong> the first section <strong>of</strong> the lowpressure<br />

turbine is improved with a higher entering mass flow rate (remember that the<br />

low-pressure turbine is designed to work in condensing mode, that is, when no steam<br />

is derived to the MSF unit).<br />

But the electricity production increases since the amount <strong>of</strong> steam <strong>and</strong> the efficiency<br />

<strong>of</strong> the low-pressure turbine have been improved. To maintain the final production, the<br />

amount <strong>of</strong> steam leaving the boiler must decrease from 156.1 to 146.6 kg/s. The<br />

redistribution <strong>of</strong> the flows inside the steam cycle was similar to the previous <strong>analysis</strong>;<br />

the low-pressure turbine produces the electricity that the high-pressure turbine<br />

cannot. This produces a negative impact when more steam is forced to flow in the<br />

low-pressure cycle (that is, passing through the condenser <strong>and</strong> not through the MSF<br />

heater). The feedwater system cools <strong>and</strong> additional fuel is required to reach the set<br />

point conditions <strong>of</strong> live steam in the boiler. Finally, the steam conditions leaving the<br />

high-pressure turbine are slightly varied (recall that the exhaust pressure is controlled<br />

by the MSF unit).<br />

Tables 7.39 <strong>and</strong> 7.40 show the F-P values for the steam power plant in design <strong>and</strong><br />

operation (when the inefficiency occurs in the recovery section <strong>of</strong> the MSF unit). The<br />

〈KP〉 matrix is shown in tables 7.41 <strong>and</strong> 7.42 for the design <strong>and</strong> inefficient case,<br />

respectively. The ∆ 〈KP〉 matrix is the key to analyze the system with this inefficiency<br />

(table 7.43). Table 7.44 contains the |I〉 matrix <strong>and</strong> the exergy cost array. The<br />

dysfunction matrix [DF] including the malfunction array MF is shown in table 7.45.<br />

Figures 7.22 <strong>and</strong> 7.23 include the impact on fuel <strong>analysis</strong> <strong>and</strong> irreversibility increase.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 237


TABLE 7.39 F-P values in design, 122 MW output power.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

238 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.40 F-P values without fouling in recovery section. MCR case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 239


TABLE 7.41 KP matrix in design. MCR case.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

240 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.42 KP matrix without fouling in recovery section. MCR case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 241


TABLE 7.43 Variation <strong>of</strong> KP without fouling in recovery section. MCR case.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

242 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.44 Irreversibility matrix without fouling in recovery section (MCR case).<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 243


TABLE 7.45 Malfunction/dysfunction matrix without fouling in recovery section (MCR case).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

244 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.46 Malfunction matrix when the fouling in recovery is varied 0,00001 m2·K/W in MCR case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 245


FIGURE 7.22 Impact on fuel <strong>analysis</strong> without fouling in RCS, MCR case.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

246 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

FIGURE 7.23 Irreversibility increase <strong>analysis</strong> <strong>of</strong> section 7.4.2.3.


<strong>Thermoeconomic</strong> diagnosis<br />

The first <strong>analysis</strong> compared the fuel impact associated with the whole plant (a fuel<br />

savings <strong>of</strong> 24.00 MW), with the fuel saved in the MSF 1 unit (26.47 MW). This means<br />

that the steam power plant is forced to work under less-efficient operating conditions.<br />

We will now explain the most significant values in the malfunction array <strong>of</strong> table<br />

7.45, relating the physical consequences to the matrix values.<br />

The deaerator component mixes <strong>and</strong> preheats the feedwater from the condenser.<br />

Irreversibility in the mixing process is lower because the mass flows entering the<br />

deaerator are lower during operation (where the live steam mass flow rate is reduced<br />

to maintain the final production) than in the design. Thermal irreversibility was lower<br />

because the cold flow entering the mixer was increased <strong>and</strong> its irreversibility was<br />

reduced by 828 kW (see table 7.45). The efficiency <strong>of</strong> the component should thereby<br />

increase, i.e., the variation <strong>of</strong> the unit exergy consumption was negative in the<br />

component (∆k = –0.085, see the ∆ 〈KP〉 table 7.43), implying an induced<br />

malfunction <strong>of</strong> –682 kW.<br />

The feedwater temperature entering the boiler was reduced because the low-pressure<br />

cycle increased its contribution to the whole system. The boiler consumed additional<br />

resources to reach the set point <strong>of</strong> the steam turbine (93 bar, 535 ºC). The increased<br />

exergy unit consumption in the boiler was ∆k = 0.004, <strong>and</strong> the malfunction associated<br />

with the component was finally 858 kW (see tables 7.45 <strong>and</strong> 7.43 respectively), with<br />

an associated fuel impact <strong>of</strong> 576 kW.<br />

The first section <strong>of</strong> the high-pressure turbine had a 1,320 kW induced malfunction as<br />

a consequence <strong>of</strong> the mathematical model. The efficiency <strong>of</strong> the Curtis blade was<br />

correlated as a function <strong>of</strong> the live steam from the boiler under different operating<br />

conditions, as in this case this amount has been decreased considerably, the<br />

isoentropic efficiency in the section decreases (<strong>and</strong> consequently the exergy <strong>and</strong><br />

entropy properties <strong>of</strong> steam leaving the section). Consequently, ∆ 〈KP〉 in table 7.43<br />

was positive (∆k = 0.0267) with a 1,320 kW malfunction <strong>and</strong> a 1,954 kW fuel impact.<br />

Surprisingly, the fourth section <strong>of</strong> the high-pressure turbine had a decreased<br />

isoentropic efficiency but a negative induced malfunction (–484 kW, see table 7.45)<br />

<strong>and</strong> 1,500 kW fuel was saved. This abnormal behavior is explained by the exhaust<br />

pressure <strong>of</strong> the high-pressure turbine which decreased with the amount <strong>of</strong> live steam,<br />

allowing the output power (produced in the section) to increase. Since the product <strong>of</strong><br />

this component is the output power (according to the thermoeconomic model), the<br />

unit exergy consumption was lower during the inefficiency, resulting in a ∆k value <strong>of</strong><br />

–0.025 (see table 7.43 with the ∆ 〈KP〉 components).<br />

1. In this case the MSF unit is the component inserted in the structure productive <strong>of</strong> the steam power plant.<br />

Exergy product <strong>of</strong> the MSF unit is kept constant.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 247


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

As mentioned above, the efficiency <strong>of</strong> the first section <strong>of</strong> the low-pressure turbine<br />

increased because the amount <strong>of</strong> entering steam increases considerably <strong>and</strong> its<br />

∆ 〈KP〉 component was negative (∆k = –0.319, table 7.43). The irreversibility <strong>of</strong> the<br />

process was reduced by 1,965 kW, with a –2,177 kW induced malfunction <strong>and</strong> 2,970<br />

kW fuel savings.<br />

The malfunction associated with the MSF was positive, although the fuel impact<br />

saved was very high in this component. The dysfunction generated by this component<br />

achieved the desired fuel savings (a large negative value). The reason for the 12,491<br />

kW induced malfunction (table 7.45, which coincides with the irreversibility increase<br />

<strong>of</strong> the process in this case) was the drastic reduction in negentropy (which was<br />

introduced in the thermoeconomic model <strong>of</strong> the steam cycles to account for the heat<br />

rejected in the condenser). This negentropy is a subproduct in the productive<br />

structure. The unit exergy consumption <strong>of</strong> the unit increased (∆k = 1.837, table 7.43).<br />

The most important dysfunctions in the boiler <strong>and</strong> condenser were caused by the<br />

components with the most important malfunctions (figure 7.24). The boiler <strong>and</strong><br />

condenser suffered dysfunctions <strong>of</strong> 1,620 kW <strong>and</strong> 752 kW from the deaerator;<br />

1,812 kW <strong>and</strong> –1,215 kW from the first section <strong>of</strong> the high pressure turbine,<br />

-- 1,391 kW <strong>and</strong> 404 kW from the fourth section <strong>of</strong> this turbine <strong>and</strong> –24.55 MW <strong>and</strong><br />

--13.68 MW from the MSF unit, respectively. The final product in the boiler was<br />

reduced by more than 10 MW to maintain the total production at a lower steam<br />

consumption (total boiler dysfunction, –22.65 MW). The condenser also increased<br />

production by 16 MW (total condenser dysfunction, –13.71 MW). The high φ<br />

coefficients promote high dysfunction since they are related to the position <strong>of</strong> the<br />

components in the productive structure <strong>of</strong> the system.<br />

Following the methodology in Chapter 6 for the diagnosis <strong>of</strong> complex systems, the<br />

DI array is the column sum <strong>of</strong> the dysfunctions. The values <strong>of</strong> the main components<br />

are described in the previous paragraph (2.43 MW for the deaerator, 634 kW for the<br />

HPT1, <strong>and</strong> –38.9 MW for the MSF unit!). The impact on fuel associated with each<br />

component (table 7.45) was obtained by adding the malfunction array MF. This is the<br />

additional fuel consumed due to the change in the operation <strong>of</strong> each unit with respect<br />

to the operating conditions <strong>and</strong> no inefficiency.<br />

Having obtained the most relevant results in the inefficiency <strong>analysis</strong>, the malfunction<br />

matrix can be used as a predictive tool to diagnose the effects <strong>of</strong> the inefficiency.<br />

Figure 7.24 shows the total impact on fuel associated with the inefficiency variation<br />

in the recovery section (the effect <strong>of</strong> fouling in the recovery section when fouling is<br />

varied). Here the malfunction matrix did not exactly predict the malfunctions because<br />

the response <strong>of</strong> the mathematical model was not perfectly linear when varying the<br />

steam to the MSF unit (under maximum production, some internal flows <strong>of</strong> the<br />

distiller are forced to keep a constant value). However, the MSF model behaved<br />

linearly at nominal production.<br />

248 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> diagnosis<br />

FIGURE 7.24 Impact on fuel depending on fouling in recovery section.<br />

-6000<br />

-12000<br />

-18000<br />

-24000<br />

0<br />

Inc. fuel consumption<br />

Since the model responded in a non-linear way to the efficiency, the most rigorous<br />

diagnosis should separate simulate each case (avoiding the malfunction matrix). The<br />

malfunction matrix in table 7.46 provides the ‘linearized’ malfunction induced in<br />

each component when the fouling in the recovery section is changed by<br />

0.00001 m 2 ·K/W. The condenser pump <strong>and</strong> low-pressure heater coefficients were<br />

again high (as are the brine heater <strong>and</strong> feed pump values), although the low product<br />

did not induce an important malfunction. As expected, the MSF coefficients were the<br />

highest <strong>and</strong> the HPT1 <strong>and</strong> HPT4 were also elevated.<br />

FIGURE 7.25 Monetary cost <strong>of</strong> electricity depending on the fouling in recovery section.<br />

0,03796<br />

0,03794<br />

0,03792<br />

0,03790<br />

0,03788<br />

0 36 9 12 15<br />

kW<br />

$/kW·h<br />

Electricity cost<br />

fouling*10-5 in RCS<br />

0 3 6 9 12 15<br />

The ‘monetary diagnosis’ (figures 7.25 <strong>and</strong> 7.26) involves the cost <strong>of</strong> electricity <strong>and</strong><br />

water as a function <strong>of</strong> recovery section fouling during maximum production in winter<br />

(which was the load requested in the example). The cost <strong>of</strong> electricity increased a bit<br />

(4×10 –6 $/kW·h) when the fouling was decreased. The malfunction <strong>analysis</strong> proved<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 249


<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

that the steam power plant decreases its global efficiency with an inefficiency in the<br />

recovery section <strong>of</strong> the MSF unit (as explained at the beginning <strong>of</strong> this section).<br />

Water cost followed the expected results, 0.0057 $/m 3 was saved with a 0.00001 m 2<br />

K/W decrease in recovery section fouling (see figure 7.26) or 120,000 $/y.<br />

FIGURE 7.26 Cost in $ per cubic meter <strong>of</strong> water when recovery section fouling is varied.<br />

1,05<br />

1,03<br />

1,01<br />

0,99<br />

0,97<br />

0,95<br />

$/m3<br />

In summary:<br />

Water cost<br />

0 36 9 12 15<br />

• The results <strong>of</strong> the inefficiency diagnosis imply that fouling in the recovery<br />

section considerably reduces the amount <strong>of</strong> steam needed to produce freshwater.<br />

The cost <strong>of</strong> water was drastically reduced (see figure 7.26) when the cleaning<br />

ball system operates in the recovery section <strong>of</strong> the MSF distillers. But a<br />

reduction in the derived steam did not imply improved plant performance (for<br />

this particular case, the electricity cost was even higher).<br />

• A consequence <strong>of</strong> this example is that the co-generation plant should operate at<br />

an optimum ratio between the steam to MSF <strong>and</strong> the live steam produced in the<br />

boiler. The installation <strong>of</strong> the cleaning ball system in the MSF distillers should<br />

be taken into account in the design in the co-generation plant, because the<br />

optimum point <strong>of</strong> the performance in the dual plant is seriously affected by the<br />

use <strong>of</strong> this system.<br />

• An inefficiency in the MSF unit provokes induced malfunctions in several<br />

components <strong>of</strong> the steam power plant (boiler, deaerator, some turbine<br />

sections…). Therefore, this type <strong>of</strong> inefficiency detected ‘downstream’ has a<br />

more global effect than an inefficiency local to a component in the steam power<br />

plant.<br />

250 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


<strong>Thermoeconomic</strong> diagnosis<br />

7.3.3 Analysis <strong>of</strong> several inefficiencies<br />

7.3.3.1 Analysis <strong>of</strong> several simultaneous inefficiencies in the steam power plant<br />

We will now consider the <strong>combined</strong> effect <strong>of</strong> several simultaneous inefficiencies in<br />

different components <strong>and</strong> the effect <strong>of</strong> induced malfunctions. This exercise reinforces<br />

the concept <strong>of</strong> local <strong>and</strong> intrinsic malfunctions. We simulated the physical effect <strong>of</strong><br />

these inefficiencies (changing main flowstreams) <strong>and</strong> describe related malfunctions,<br />

dysfunctions, <strong>and</strong> additional fuel consumption in the steam power plant (the direct<br />

problem).<br />

The analyzed inefficiencies were:<br />

• TTD in high-pressure heater no. 1 increases 5 ºC<br />

• Isoentropic efficiency <strong>of</strong> the feed pump decreases 10%.<br />

• Isoentropic efficiency <strong>of</strong> the first section <strong>of</strong> the high-pressure turbine<br />

decreases 5%.<br />

• Isoentropic efficiency <strong>of</strong> the first section <strong>of</strong> the low-pressure turbine decreases<br />

15%.<br />

• Isoentropic efficiency <strong>of</strong> the fourth section <strong>of</strong> the high-pressure turbine<br />

decreases 10%.<br />

If the TTD <strong>of</strong> the HPH1 increases, the feedwater leaves the heater at a lower<br />

temperature <strong>and</strong> the turbine extraction temperature increases. If the heater does not<br />

need to preheat the feedwater the same amount, the extraction mass flow should be<br />

reduced. The boiler is also affected because feedwater enters the economizers at a<br />

lower-than-design temperature.<br />

The mechanical irreversibility <strong>of</strong> the feed pump increases if the isoentropic efficiency<br />

is lower than expected. The pump responds by consuming more power <strong>and</strong> the<br />

feedwater temperature increases.<br />

The exhaust conditions <strong>of</strong> the high <strong>and</strong> low-pressure turbine are more or less<br />

maintained with the MSF unit <strong>and</strong> ambient conditions. When the isoentropic<br />

efficiency <strong>of</strong> several sections <strong>of</strong> the steam turbine decreases, the steam conditions are<br />

not significantly affected by inefficiencies in other sections. The output power in each<br />

inefficient section is not enough to maintain final production but other sections cannot<br />

produce this extra power since their efficiency was maintained constant. Thus,<br />

although the system dem<strong>and</strong>s more live steam, the efficiency <strong>of</strong> the boiler does not<br />

necessarily decrease.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 251


TABLE 7.47 F-P values in design, 122 MW output power.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

252 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.48 F-P values with inefficiencies in five components (MCR case).<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 253


TABLE 7.49 KP matrix in design (MCR Case).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

254 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.50 KP matrix with several inefficiencies in MCR case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 255


TABLE 7.51 Variation <strong>of</strong> KP matrix with several inefficiencies in MCR case.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

256 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE 7.52 Irreversibility matrix with five inefficiencies in power plant (MCR case).<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 257


TABLE 7.53 Malfunction/dysfunction matrix with five inefficiencies in MCR case.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

258 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE 7.27 Impact on fuel <strong>analysis</strong> in section 7.3.3.1.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

FIGURE 7.28 Irreversibility increase in section 7.3.3.1.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 259


CHAPTER 8<br />

Synthesis, contributions <strong>and</strong><br />

perspectives<br />

8.1 Synthesis<br />

This Ph. D. Thesis brings together three topics that have never been thoroughly<br />

interrelated.<br />

• Desalination processes.<br />

Water scarcity is a serious problem for humanity now <strong>and</strong> in the future. Water<br />

resources are being depleted by excessive consumption <strong>and</strong> polluted by human<br />

development. Fortunately, the problem can be solved by desalting seawater or<br />

reusing wastewater. Chapter 1 describes the current situation in arid countries <strong>and</strong><br />

how to solve some water shortages. Chapter 2 summarizes the most common<br />

methods to produce freshwater for human consumption.<br />

• Energy consumed in desalination.<br />

The detailed description <strong>of</strong> desalination processes in Chapter 2 including the<br />

consumption <strong>and</strong> energy producing process in desalination. It is very energy<br />

intensive <strong>and</strong> should not be isolated from energy production processes.<br />

Desalination designers normally present the energy consumption <strong>of</strong> different<br />

3<br />

desalination processes in terms <strong>of</strong> electrical consumption (kW·h/m ) even if they<br />

consume thermal energy. The current trend is to separate the two processes. The<br />

existence <strong>of</strong> big companies that only produce electricity or only water widens the<br />

gap between desalination <strong>and</strong> energy communities. This thesis demonstrates that<br />

energy <strong>and</strong> water suppliers interact in a co-generation installation to provide both<br />

products <strong>and</strong> that both systems should not be analyzed separately.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


324<br />

Synthesis, contributions <strong>and</strong> perspectives<br />

• <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> the most common desalting <strong>and</strong> power installations.<br />

We used thermoeconomic techniques normally applied to power plants. In this<br />

way, we took advantage <strong>of</strong> everything that thermoeconomics provides to obtain<br />

an in depth knowledge <strong>of</strong> a very complex system. The energy supplier was also<br />

analyzed since the desalting plant is coupled with the power plant. We analyzed<br />

one <strong>of</strong> the most common processes used in arid regions with important water<br />

scarcity problems: multi-stage flash desalting plants that use fossil fuels to also<br />

produce electricity with the help <strong>of</strong> a conventional power plant. The<br />

thermoeconomic <strong>analysis</strong> was also applied to a steam power plant providing<br />

steam to the MSF unit.<br />

The main contribution <strong>of</strong> the thesis is contained in Chapter 7. The thermoeconomic<br />

<strong>analysis</strong> <strong>of</strong> the dual plant included cost <strong>analysis</strong>, diagnosis, <strong>and</strong> optimization. The<br />

results <strong>of</strong> the different thermoeconomic techniques applied in each case are as<br />

follows:<br />

1. The cost <strong>analysis</strong> is very useful to find the enormous possibilities <strong>of</strong> energy<br />

savings under different configurations <strong>of</strong> the co-generation plant. A detailed<br />

<strong>analysis</strong> <strong>of</strong> the internal costs pin-points the component responsible for<br />

irreversibilities. New processes can also be <strong>combined</strong> to produce minimum water<br />

<strong>and</strong> electricity costs.<br />

2. Plant diagnosis helps to elucidate component interaction. The different<br />

relationships <strong>and</strong> effects <strong>of</strong> component inefficiencies on other subsystems can be<br />

successfully quantified by considering both plants together. The interaction can<br />

also be separated by varying component efficiency (malfunction) <strong>and</strong> the<br />

subsequent additional component production (dysfunction). This thesis includes<br />

one example <strong>of</strong> a thermodynamically isolated (power plant) <strong>and</strong> non-isolated<br />

(MSF unit) system. However, the diagnosis cannot be used as a predictive tool in<br />

the control systems because the theory cannot yet recognize the origin <strong>of</strong> the<br />

inefficiencies.<br />

3. Local optimization optimizes the operating conditions by calculating the<br />

minimum product cost <strong>of</strong> each plant unit. It is very valuable to design new cogeneration<br />

plants or to readapt existing ones.<br />

4. Product cost <strong>and</strong> price must be calculated from their origin. Cost is the resources<br />

consumed to produce something <strong>and</strong> price is the value obtained when this product<br />

is sold. Benefit is the difference between both concepts. Once the price is known,<br />

the cost must be minimized to obtain the maximum benefit (plant operating<br />

conditions <strong>of</strong> the plant can be changed intentionally depending on dem<strong>and</strong> <strong>and</strong><br />

price).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Main contributions<br />

8.2 Main contributions<br />

This Ph. D. Thesis applies the most recent thermoeconomic techniques (normally<br />

only applied to power generation systems) to a power <strong>and</strong> desalination plant. The<br />

main contributions <strong>of</strong> the thesis are listed here:<br />

8.2.1 Simulator <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant<br />

A simulator was used to provide the thermodynamic states <strong>of</strong> desalination <strong>and</strong> power<br />

generation process. Thermal desalination processes have been simulated by chemical<br />

engineers (Jernqvist, Jernqvist <strong>and</strong> Aly, 1999; Ettouney <strong>and</strong> El-Dessouky, 1999), but<br />

the steam producing system is not considered. The two processes were separately<br />

introduced in the simulator to independently analyze each process. A <strong>combined</strong> state<br />

can be modelled by introducing the same quantity <strong>of</strong> steam sent to the MSF unit. The<br />

simulator was validated using performance data cases <strong>and</strong> real operating data for the<br />

dual-plant with a MSF unit <strong>and</strong> a steam power plant. It can model the effect <strong>of</strong><br />

inefficiencies in the two systems for diagnosis.<br />

The mathematical model applied under different operating modes accurately<br />

reproduces (for engineering purposes) the real state <strong>of</strong> the plant, despite the scarcity<br />

<strong>of</strong> data for each operating mode. The most difficult case is when the amount <strong>of</strong> steam<br />

entering the low-pressure turbine is so low that the system has to consume<br />

mechanical energy to move the blades. In this case, the input conditions <strong>of</strong> the<br />

mathematical model have to be continuously restricted in order to preserve the<br />

stability <strong>of</strong> the model. The mathematical models <strong>of</strong> the MSF <strong>and</strong> steam turbine power<br />

plant were solved using a solution algorithm that simultaneously h<strong>and</strong>les the whole<br />

set <strong>of</strong> model equations. The packages containing the sequential scheme to solve the<br />

flowsheeting <strong>of</strong> a plant are discarded here, although this threatens model stability<br />

under different operating conditions.<br />

8.2.2 State <strong>of</strong> the art in <strong>Thermoeconomic</strong>s<br />

An effort was made in Chapter 6 to review <strong>and</strong> summarize <strong>Thermoeconomic</strong><br />

methodologies. The Structural Theory was finally adopted to explain the concepts,<br />

procedures <strong>and</strong> applications <strong>of</strong> these techniques, including the matrix formulation<br />

<strong>and</strong> new terms like induced malfunction, intrinsic malfunction <strong>and</strong> dysfunction. The<br />

thermoeconomic <strong>analysis</strong> <strong>of</strong> the dual plant was based on this theory <strong>and</strong> its latest<br />

improvements.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

325


326<br />

Synthesis, contributions <strong>and</strong> perspectives<br />

8.2.3 F-P definition for a MSF unit<br />

The F-P definition <strong>of</strong> the thermoeconomic <strong>analysis</strong> <strong>of</strong> the MSF unit (see section<br />

7.1.3.2) highlights the cost <strong>of</strong> water production in the recovery <strong>and</strong> reject section,<br />

taking into account the thermodynamic processes in the plant. Several F-P definitions<br />

solved the model but none gave appropriate costs <strong>of</strong> device functionality nor for the<br />

main flow degradation in the MSF plant.<br />

8.2.4 Cost <strong>analysis</strong> <strong>of</strong> a dual-plant<br />

A detailed cost <strong>analysis</strong> <strong>of</strong> the power <strong>and</strong> desalination plant was carried out under<br />

different operating modes (see section 7.2). The physical (or formation) costs <strong>of</strong> the<br />

main components were calculated. Exergy operating costs are available for each<br />

component as well as the thermoeconomic costs <strong>of</strong> water <strong>and</strong> electricity. These latter<br />

costs were successfully compared with other methodologies (EL-Nashar, 1999;<br />

Kronenberg et al., 1999) that do not use the thermoeconomic model <strong>and</strong> provide<br />

much less information.<br />

8.2.5 Diagnosis <strong>of</strong> a complex system<br />

The thermoeconomic diagnosis in section 7.3 was based on Structural Theory <strong>and</strong><br />

Symbolic <strong>Thermoeconomic</strong>s (Torres et al., 1999). This is the first time that a<br />

malfunction/dysfunction <strong>analysis</strong> is applied to a complex energy system (26<br />

components <strong>and</strong> 4 junctions for the power plant <strong>and</strong> 11 components <strong>and</strong> 7 junctions<br />

for the MSF plant). Usually the matrix formulation is only used to study simpler<br />

systems like the gas turbine co-generation plant in Chapter 6. The malfunction/<br />

dysfunction table provides a lot <strong>of</strong> information that should be carefully analyzed<br />

when an inefficiency is simulated in the plant (exergy costs, impact on fuel,<br />

irreversibility increase in each component...). The relationships between components<br />

are rapidly found with in terms <strong>of</strong> efficiency variation (intrinsic or induced<br />

malfunction) or additional production (dysfunction). This method does not find the<br />

nature <strong>of</strong> the malfunction. Whether it is intrinsic or induced depends on user<br />

knowledge.<br />

The symbolic notation <strong>and</strong> Structural Theory also helps to formulate the malfunction<br />

matrix to find the quantity <strong>of</strong> additional resources consumed due to an inefficiency<br />

(without using the simulator). This matrix is used when the system responds linearly<br />

to the applied inefficiencies. If the inefficiency is local to the component, individual<br />

matrices <strong>of</strong> different inefficiencies may be added to make one large malfunction<br />

matrix with the same effect.<br />

To date, most analyzed systems demonstrate additivity <strong>of</strong> diagnosis: several<br />

inefficiencies can be disaggregated. However, the MSF did not fulfil this requirement<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Perspectives<br />

in the diagnosis <strong>of</strong> complex systems. As seen above, this fulfilment strongly depends<br />

on the physical structure <strong>of</strong> the system.<br />

8.2.6 Local optimization <strong>of</strong> the steam power plant<br />

Local optimization <strong>of</strong> the main components <strong>of</strong> the steam power plant also provides a<br />

global minimum final cost <strong>of</strong> electricity <strong>and</strong> water. Global optimization <strong>of</strong> the steam<br />

power plant (based on local optimization, see section 7.5) has never been applied to a<br />

set <strong>of</strong> 14 free-design variables that govern plant behavior. Local optimization can be<br />

applied to the steam power plant because it is thermodynamically isolated, i.e. local<br />

perturbations only affect one component (demonstrated in the diagnosis <strong>of</strong> the steam<br />

power plant).<br />

8.2.7 Cost, price <strong>and</strong> benefit<br />

Finally, a new methodology is included to assign product cost, price <strong>and</strong> benefit using<br />

examples to demonstrate that cost <strong>and</strong> price are independent. The main objective <strong>of</strong><br />

an investment is to obtain maximum benefit, which does not always imply minimum<br />

cost.<br />

8.3 Perspectives<br />

8.3.1 Improving existing plants. Process integration<br />

One <strong>of</strong> the immediate consequences <strong>of</strong> this work is to increase the ways existing<br />

plants may reduce energy consumption. After analyzing one <strong>of</strong> the most developed<br />

methods to produce freshwater <strong>and</strong> electricity, some areas were found lacking. Our<br />

suggestions include:<br />

• Promoting the <strong>simulation</strong> <strong>of</strong> both processes (water <strong>and</strong> energy production)<br />

integrated in specific simulators.<br />

• Applying our methodology to other desalination processes. Our objective was to<br />

find the most efficient process at the lowest energy consumption, the best way to<br />

produce both energy <strong>and</strong> water <strong>and</strong> not contribute to fossil fuel depletion, air<br />

pollution <strong>and</strong> climate change. The importance <strong>of</strong> hybrid configurations, i.e. the<br />

integration <strong>of</strong> other processes to produce energy (wind, solar, tides, even nuclear)<br />

<strong>and</strong> water (MSF/RO or MED/RO units, heat absorption pumps) will possibly be<br />

the trend in the next decades. 'Building-block’ s<strong>of</strong>tware will be required to<br />

thermoeconomically analyze any process producing water or electricity.<br />

• <strong>Thermoeconomic</strong>s only considers the costs <strong>of</strong> operation, installation <strong>and</strong><br />

maintenance, but processes also involve other costs that should be taken into<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

327


328<br />

Synthesis, contributions <strong>and</strong> perspectives<br />

account including environmental (pollution, brine discharges...), costs <strong>of</strong><br />

producing materials, biological costs, building costs, etc. All these are less<br />

developed than the costs evaluated in the thermoeconomic <strong>analysis</strong>. Since<br />

thermoeconomic techniques can consider any type <strong>of</strong> costs, they should be<br />

introduced in the global theory when they are more or less available.<br />

8.3.2 Improvements in thermoeconomic diagnosis<br />

Cost <strong>analysis</strong> provides a lot <strong>of</strong> information about how processes degrade <strong>and</strong> the<br />

energy quality <strong>of</strong> fluids in a plant. It is very useful to quantify the efficiency <strong>of</strong> plant<br />

processes. Diagnosis is directly oriented to an on-line implementation in the control<br />

system. In this regard, a big effort is needed to improve thermoeconomic techniques<br />

related to plant diagnosis (the ‘ inverse problem’)<br />

when the data acquisition system<br />

(DAS) finds deviations from the target conditions for each operating mode <strong>and</strong> load.<br />

The diagnosis should detect the inefficiency from the data collected by the DAS to<br />

take corrective actions. New communication technologies (Internet) allow remote<br />

control <strong>of</strong> the on-line implementation <strong>of</strong> system diagnosis, so plant managers can also<br />

see the benefits <strong>of</strong> the implementation. The on-line system can also be installed<br />

higher up in the control system, i.e. it can be used for all units. The units respond as a<br />

whole unit when a deviation is detected. Regarding maximizing benefit, the higher<br />

level <strong>of</strong> hierarchy can help decide the most pr<strong>of</strong>itable configuration <strong>and</strong> the<br />

possibility <strong>of</strong> connecting the hybrid systems installed in the plants for additional<br />

water or energy in peak or low-dem<strong>and</strong> periods.<br />

Some previous steps may be needed to solve the inverse problem <strong>of</strong> the<br />

thermoeconomic diagnosis:<br />

• Analyze the problem <strong>of</strong> ‘noise’ provoked by the real boundary conditions in an<br />

installation: set points, ambient conditions, fuel quality, different loads <strong>and</strong><br />

operating modes. We should consider the way to isolate the system from these<br />

boundary conditions or their effects. Once the problem is solved, the diagnosis<br />

finds the real causes <strong>of</strong> the deviations.<br />

• <strong>Thermoeconomic</strong>s should be used to investigate the development <strong>of</strong> new<br />

techniques to study component interdependence during induced malfunctions in a<br />

complex system. The non-additivity <strong>of</strong> the diagnosis in these interrelated systems<br />

opens an interesting new line <strong>of</strong> investigation.<br />

• Clarifying the application <strong>and</strong> interpretation <strong>of</strong> dysfunctions generated in/by plant<br />

components. We could consider performing the <strong>analysis</strong> under a constant final<br />

production (<strong>of</strong> the complex system), depending on the finality <strong>of</strong> the <strong>analysis</strong>.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Perspectives<br />

8.3.3 Integrating attitudes<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> provides enormous amounts <strong>of</strong> information about plant<br />

functioning <strong>and</strong> possible savings. This information should be clearly integrated in a<br />

vertical structure, i.e. a different kind <strong>of</strong> information should go to each level in the<br />

plant staff hierarchy. For example, if we divide the organization <strong>of</strong> the plant in three<br />

levels, we have:<br />

Operator level<br />

The information derived from the diagnosis is the most important at this level. The<br />

physical <strong>and</strong> economic effects <strong>of</strong> the inefficiencies <strong>and</strong> the control strategies (security<br />

versus economy) are the main issues for operators.<br />

Technician level<br />

This field includes optimizing existing systems <strong>and</strong> investigating <strong>and</strong> developing<br />

more efficient systems, <strong>and</strong> new control systems to h<strong>and</strong>le inefficiencies.<br />

Managers level<br />

Cost <strong>analysis</strong> must be the main tool used by the plant managers since they manage<br />

the whole plant (assuming there are many units per plant). Cost, price <strong>and</strong> benefit<br />

must be clearly differentiated at this level.<br />

Training seminars are necessary for all levels to inform staff about the<br />

“thermoeconomic culture” <strong>and</strong> its benefits for humanity.<br />

8.3.4 Sustainable desalination<br />

Desalination is one <strong>of</strong> the most promising means <strong>of</strong> producing drinkable water with a<br />

low impact on the environment. The tendency <strong>of</strong> the desalination scientific<br />

community is to reduce energy consumption <strong>and</strong> substitute primary energy sources<br />

by renewable sources on a large scale. This tendency should be followed in all areas<br />

that influence our future. A more global <strong>analysis</strong>, like the Life Cycle Analysis (LCA),<br />

including additional aspects (residues, product use, materials, etc) is also necessary to<br />

provide an overall perspective <strong>of</strong> desalination processes.<br />

Research on desalination using solar energy for existing or new methods, should be<br />

encouraged. As solar technology develops, the cost <strong>of</strong> producing water (on a large<br />

scale) will decrease as will the strong dependence on energy.<br />

Promoting the installation <strong>of</strong> simple devices to provide water in acceptable conditions<br />

at a very low (or zero) cost in non-developed/isolated areas (Africa, India), is another<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

329


330<br />

Synthesis, contributions <strong>and</strong> perspectives<br />

means <strong>of</strong> redistributing world resources <strong>and</strong> promoting a more equal development in<br />

the world community.<br />

8.3.5 Promote energy <strong>and</strong> water interactions<br />

Water <strong>and</strong> energy are both limited resources, vital to the quality <strong>of</strong> the human life.<br />

The rapidly growing human population increases the dem<strong>and</strong> for these resources<br />

every day. Several international organizations are dedicated to energy <strong>and</strong> several<br />

others to water, but there is a marked lack <strong>of</strong> attention to <strong>combined</strong> water <strong>and</strong> energy<br />

issues.<br />

This Ph. D. Thesis demonstrates that energy <strong>and</strong> water cannot be studied separately.<br />

A multi-disciplinary group <strong>of</strong> water <strong>and</strong> energy specialists has been formed<br />

(International Study Group for Water <strong>and</strong> Energy Systems (ISGWES), settled at the<br />

University <strong>of</strong> Zaragoza) to promote the interchange <strong>of</strong> ideas, scientific knowledge<br />

<strong>and</strong> sustainable development <strong>of</strong> water <strong>and</strong> energy systems. Some <strong>of</strong> the investigation<br />

lines commented above will be promoted by ISGWES.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


ANNEX 1<br />

<strong>Thermoeconomic</strong> diagnosis<br />

The thermoeconomic diagnosis <strong>of</strong> the dual plant in section 7.3 considered several<br />

inefficiencies described at the beginning <strong>of</strong> the chapter. Each inefficiency requires<br />

many tables <strong>and</strong> figures, all <strong>of</strong> which are included in this annex. Thus, this annex is<br />

an overall view <strong>of</strong> the effects provoked by one or more inefficiencies in the power<br />

<strong>and</strong> desalination plant.<br />

The following individual inefficiencies (described but not analyzed in section 7.3)<br />

were applied:<br />

• Inefficiency in the HPH1: variation in terminal temperature difference <strong>of</strong> heater.<br />

• Inefficiency in the feed pump: reduced efficiency.<br />

• Inefficiency in the high-pressure turbine: efficiency <strong>analysis</strong> in the first section.<br />

• Inefficiency in the low-pressure turbine: efficiency variation in first section.<br />

• Inefficiency in the recovery section: effect <strong>of</strong> reduced fouling in MSF.<br />

• Inefficiency in the reject section: effect <strong>of</strong> the fouling factor.<br />

The <strong>analysis</strong> was performed under different operating conditions but is only<br />

presented for one load, the MCR case for the power plant <strong>and</strong> NTOS performance<br />

case for the MSF unit. The effect <strong>of</strong> the different loads in the two systems is<br />

summarized in sections 7.3.4 <strong>and</strong> 7.3.5.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


332<br />

<strong>Thermoeconomic</strong> diagnosis<br />

A1.1 Effect <strong>of</strong> an inefficiency in the high-pressure heater<br />

no.1 (HPH1)<br />

The TTD <strong>of</strong> the HPH1 was varied to analyze the effect on the steam power plant.<br />

The heater TTD is the difference between the temperature <strong>of</strong> saturated vapor<br />

extracted from the turbine <strong>and</strong> the feedwater leaving the heater. Since the conditions<br />

<strong>of</strong> the steam extracted in the turbine are maintained, a higher TTD implies a poorer<br />

heat transfer inside the heater tubes. The feedwater therefore leaves the heater at a<br />

lower temperature than expected. Consequently, the extraction mass flow to this<br />

heater decreases <strong>and</strong> the boiler produces less live steam. Although the live steam<br />

needed for the electricity dem<strong>and</strong> is reduced, the boiler heats the feedwater from a<br />

lower temperature <strong>and</strong> natural gas consumption increases. An excessive change in<br />

heater TTD may also sharply vary the levels inside the heater, leading to dangerous<br />

problems or even drains in the heaters. The consequences are very difficult to<br />

evaluate with conventional component <strong>analysis</strong> since the model does not incorporate<br />

the security system layout <strong>of</strong> the power plants.<br />

The mathematical explanation <strong>of</strong> varying TTD involves the malfunction <strong>and</strong><br />

dysfunction matrices detailed in section 7.3. Tables A1.1 <strong>and</strong> A1.2 include the F-P<br />

definition <strong>of</strong> the steam power plant in design <strong>and</strong> operation with an inefficiency in<br />

the HPH1: the TTD increases 5 ºC. The output power used was 122 MW, but other<br />

examples were at 60, 90 <strong>and</strong> 140 MW, corresponding to the parallel mode,<br />

extraction mode with partial load <strong>and</strong> condensing mode, respectively (section 7.3.4).<br />

These are the most important operating modes (the most operating hours per year) in<br />

the power <strong>and</strong> desalination plant. Tables A1.3 <strong>and</strong> A1.4 include the 〈 KP〉<br />

tables<br />

corresponding to the design <strong>and</strong> inefficient operation, <strong>and</strong> table A1.5 is the ∆ 〈 KP〉<br />

matrix. Table A1.6 contains the φ coefficients <strong>of</strong> the irreversibility matrix | I〉<br />

with<br />

exergy cost <strong>of</strong> components. Finally, table A1.7 is the malfunction/dysfunction table<br />

built using table A1.6. Table A1.8 is the malfunction matrix when we vary the TTD<br />

<strong>of</strong> the HPH1 1 ºC. Figures A1.1 <strong>and</strong> A1.2 show the impact on fuel <strong>analysis</strong> <strong>and</strong> the<br />

irreversibility increase per component.<br />

The highest malfunctions in table A1.7 corresponded to the boiler, HPH1 (the<br />

inefficient component) <strong>and</strong> HPT1. The rest <strong>of</strong> the components were within simulator<br />

accuracy (< 100 kW). Varying the heater TTD only affected the components<br />

interacting with the heater. This inefficiency did not induce malfunctions in other<br />

components.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.1 F-P values in design (MCR case).<br />

Effect <strong>of</strong> an inefficiency in the high-pressure heater no.1 (HPH1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

333


TABLE A1.2 F-P values in operation with 5º C TTD respect to design .<br />

334<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.3 KP matrix in design (MCR case).<br />

Effect <strong>of</strong> an inefficiency in the high-pressure heater no.1 (HPH1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

335


TABLE A1.4 KP matrix with inefficiency in HPH1 (MCR case).<br />

336<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.5 Variation <strong>of</strong> KP matrix when TTD in the HPH1 is 5 ºC higher than the expected.<br />

Effect <strong>of</strong> an inefficiency in the high-pressure heater no.1 (HPH1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

337


TABLE A1.6 Irreversibility matrix with the inefficiency in HPH1.<br />

338<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.7 Malfunction/Dysfunction matrix when the TTD in HPH1 is 5º C higher.<br />

Effect <strong>of</strong> an inefficiency in the high-pressure heater no.1 (HPH1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

339


TABLE A1.8 Malfunction matrix when TTD in HPH1 is varied 1 ºC<br />

340<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE A1.1 Impact on fuel <strong>analysis</strong> with an inefficiency in HPH1.<br />

Effect <strong>of</strong> an inefficiency in the high-pressure heater no.1 (HPH1)<br />

FIGURE A1.2 Irreversibility <strong>analysis</strong> when the TTD in HPH1 is increased 5 ºC.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

341


342<br />

<strong>Thermoeconomic</strong> diagnosis<br />

In the simulated intrinsic malfunction, increasing the TTD <strong>of</strong> the HPH1 by 5 ºC<br />

(which could be interpreted as a problem in the heat transfer mechanism <strong>of</strong> the<br />

heater), decreases more than expected the feedwater temperature leaving the heater.<br />

The extraction flow to this heater also decreases (to meet the energy balance <strong>of</strong> the<br />

heater). In any case, the inefficiency increases the irreversibility in the heater<br />

( ∆I<br />

= 16.5 kW, see table A1.7) when the temperature difference in the water tubes<br />

increases. The first effect (a lower heating process in the heater) is more important<br />

than the second (a lower extraction flow). Unit exergy consumption varied by<br />

∆k<br />

= 0.020 (see table A1.5). The result <strong>of</strong> the inefficiency was a 223.6 kW intrinsic<br />

malfunction (or an associated 272 kW impact on fuel).<br />

The effect induced in the boiler is clear: if the feedwater leaves the heater at a lower<br />

temperature, the boiler consumes additional fuel to maintain the live steam<br />

conditions (which are fixed in the simulator <strong>and</strong> the real plant). The ∆ 〈 KP〉<br />

component, i.e. the variation <strong>of</strong> component unit exergy consumption (table A1.5)<br />

was ∆k<br />

= 0.005. As the boiler product was very high (the heat transferred to the<br />

feedwater was about 210 MW), the malfunction was 1,179 kW with an associated<br />

830 kW impact on fuel. The total impact on fuel associated with this inefficiency<br />

was 1,048 kW. In this case, the malfunction induced in the boiler was more<br />

important than the intrinsic malfunction in the heater.<br />

The amount <strong>of</strong> steam flowing in HPT1 was lower than in design, although this effect<br />

disappears when HPH1 extraction was reduced. The steam flowing through the<br />

second section <strong>of</strong> the HPH is maintained. The energy production in this section is<br />

maintained, but the efficiency in this section is lightly decreased (the efficiency <strong>of</strong><br />

the Curtis blade is higher as the live steam flow grows), then the variation <strong>of</strong> the unit<br />

exergy consumption <strong>of</strong> the section is ∆k<br />

= 0.0026, then we have a little malfunction<br />

induced in this section <strong>of</strong> 130 kW, <strong>and</strong> an impact on fuel associated <strong>of</strong> 209 kW.<br />

The dysfunctions due to the HPH1 inefficiency emphasizes the results <strong>of</strong> other<br />

inefficiencies in the steam power plant: only the boiler <strong>and</strong> the condenser suffer<br />

dysfunctions generated by component malfunctions (HPH1, boiler or HPT1). The<br />

dysfunction generated in the boiler was positive. The φ coefficients were positive but<br />

negative for the condenser, provoking a negative dysfunction. The junction J2<br />

produced a –393 kW dysfunction in the boiler associated with its exergy unit<br />

consumption variation. This variation is explained by the productive structure <strong>of</strong> the<br />

steam power plant (see figure 7.5): feedwater heated in the boiler in one <strong>of</strong> the inlets<br />

<strong>of</strong> that junction.<br />

The power plant model varied linearly to a one degree change in the heater TTD<br />

when comparing the amount <strong>of</strong> fuel saved. Figure A1.3 shows this effect for<br />

electricity production in the extraction mode <strong>of</strong> the example (122 MW). The effect is<br />

non-linear when TTD is negative <strong>and</strong> close to zero (the design TTD for this<br />

operating condition is –1.7 ºC). The <strong>analysis</strong> could be performed avoiding this range<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE A1.3<br />

FIGURE A1.4<br />

Effect <strong>of</strong> an inefficiency in the high-pressure heater no.1 (HPH1)<br />

<strong>of</strong> temperature differences since the impact on fuel associated with the whole plant<br />

can be less than 200 kW <strong>and</strong> the mathematical model cannot diagnose the<br />

inefficiency with less than 100 kW accuracy.<br />

Impact on fuel associated with a variation in the TTD <strong>of</strong> HPH1. 122 MW power plant production.<br />

Inc. fuel consumption<br />

1200<br />

kW<br />

800<br />

400<br />

-5 -4 -3 -2 -1 0 1 2 3 4 5<br />

-400<br />

-800<br />

-1200<br />

0<br />

In any case, the observed trend could be used to apply the malfunction matrix to this<br />

inefficiency. Some matrix components had large values. The condenser pump <strong>and</strong><br />

low-pressure heater No.2 were high due to the behavior <strong>of</strong> the mathematical model<br />

at the condenser exit area (see section 4, mathematical model <strong>of</strong> the power plant).<br />

The feed pump <strong>and</strong> deaerator also had considerable values due to the decrease in<br />

feed water flow in the high-pressure zone (provoked by the HPH1 inefficiency).<br />

Cost <strong>of</strong> electricity when varying TTD in HPH1 (MCR performance case).<br />

Electricity cost<br />

0,03790<br />

0,03785<br />

0,03780<br />

0,03775<br />

$/kwh<br />

TTD (º C) in HPH1<br />

-5 -4 -3 -2 -1 0 1 2 3 4 5<br />

TTD (º C) in HPH1<br />

The effect on the cost <strong>of</strong> electricity <strong>and</strong> water was not as important as the impact on<br />

fuel associated with the inefficiency in the turbine section. It implied an additional<br />

3<br />

0.000009 $/kW·h in electricity <strong>and</strong> 0.00017 $/m in freshwater per degree Celsius in<br />

the TTD <strong>of</strong> the HPH1. This could save 9,600 $ <strong>and</strong> 3,570 $ if the plant were<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

343


FIGURE A1.5<br />

344<br />

<strong>Thermoeconomic</strong> diagnosis<br />

operating yearlong at these loads. Figures A1.4 <strong>and</strong> A1.5 refer to this assumption,<br />

emphasizing the linearity <strong>of</strong> the model (except from –2 to 0 ºC).<br />

Cost <strong>of</strong> water when varying TTD in the first HPH (MCR performance case).<br />

In summary:<br />

Water cost<br />

1,2730<br />

$/m3<br />

1,2725<br />

1,2720<br />

1,2715<br />

1,2710<br />

-5 -4 -3 -2 -1 0 1 2 3 4 5<br />

TTD (º C) in HPH1<br />

• the heater TTD affects heater behavior <strong>and</strong> components receiving feedwater<br />

heated by the inefficient heater (the boiler). The inefficiency did not result only<br />

local to its component, <strong>and</strong> the associated malfunctions were higher in other<br />

components than the intrinsic one. The rest <strong>of</strong> the components were not<br />

considerably affected compared to an inefficiency in the steam turbine sections.<br />

• The impact on fuel associated with the additional cost <strong>of</strong> water or energy due to<br />

the inefficiency was not important when compared with other inefficiencies (the<br />

total saving <strong>of</strong> 14,000 $/y in both products could be obtained by decreasing the<br />

TTD <strong>of</strong> the HPH1 by 1 ºC). This only refers to the range where the model<br />

responds linearly to TTD variation (the variational <strong>analysis</strong> was assumed to be<br />

linear). If the TTD is abnormally high, an excess heater level or excessive<br />

heating in the economizers can lead to extreme induced malfunctions that can<br />

not be calculated in the diagnosis (Valero, Torres <strong>and</strong> Lerch, 1999). Therefore,<br />

heater TTD should be carefully controlled. A by-pass in one <strong>of</strong> the HPHs is a<br />

very good example <strong>of</strong> heater inefficiency, but it is very difficult to simulate. The<br />

model needs to be modified considerably to consider this inefficiency.<br />

• The results <strong>of</strong> the HPH1 inefficiency could be extrapolated to HPH2, taking into<br />

account the amount <strong>of</strong> heat transferred in the two heaters (usually the HPH2 uses<br />

less steam to heat the feedwater). The effect on the boiler should also be<br />

reduced.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Effect <strong>of</strong> feed pump isoentropic efficiency<br />

A1.2 Effect <strong>of</strong> feed pump isoentropic efficiency<br />

The feed pump pressurizes the feedwater before it enters the boiler. An inefficiency<br />

inside the pump mechanism (assuming that the pump can supply the specified<br />

pressure) only slightly increases feedwater temperature since the temperature rise in<br />

pumping a liquid is also low. Therefore, this inefficiency should not induce<br />

important malfunctions in other components. The most important consequence is<br />

the significant increase in feed pump power consumption. Additional live steam is<br />

required to maintain the net output power.<br />

If the feed pump is coupled with an auxiliary turbine providing energy, an<br />

inefficiency should affect other components because an abnormally functioning<br />

auxiliary turbine would redistribute the flows in the steam/water cycle <strong>of</strong> the power<br />

plant.<br />

Feed pump behavior can be studied considering an isoentropic efficiency, a variable<br />

that appears in our mathematical model. Pump efficiency decreased 12% with<br />

respect to its characteristic curve at 122 MW (MCR performance case). The<br />

inefficiency was also analyzed under different operating conditions (see<br />

section 7.3.4). Tables A1.9 <strong>and</strong> A1.10 show the F-P values for design <strong>and</strong> operating<br />

conditions. The 〈 KP〉<br />

matrices are written dividing fuels <strong>and</strong> the product <strong>of</strong> each<br />

component (tables A1.11 <strong>and</strong> A1.12). After these matrices are built, the ∆ 〈 KP〉<br />

matrix <strong>and</strong> irreversibility matrix I are immediately processed, containing the unit<br />

exergy costs <strong>of</strong> the components (tables A1.13 <strong>and</strong> A1.14). The malfunction/<br />

dysfunction matrix with the dysfunction coefficients is included in table A1.15. The<br />

malfunction matrix with the extra consumption when the pump isoentropic<br />

efficiency increases 1% is finally included (table A1.16). Figures A1.6 <strong>and</strong> A1.7<br />

show the impact on fuel <strong>and</strong> the irreversibility increase in all components for this<br />

simulated inefficiency.<br />

We will now explain the physical <strong>analysis</strong> using results from the inefficiency<br />

diagnosis. The malfunction array demonstrates that the feed pump does not induce<br />

any malfunction in the rest <strong>of</strong> the components. Only the boiler <strong>and</strong> the inefficient<br />

component have a malfunction greater than 30 kW. The mechanical irreversibility<br />

increases when the pump has serious problems to reach the dem<strong>and</strong>ed pressure.<br />

These internal frictions also increase the temperature <strong>of</strong> the pressurized feedwater<br />

leaving the pump. So, the thermal irreversibility also appears in the inefficiency <strong>and</strong><br />

the final reversibility increase was ∆I<br />

= 475 kW (see table A1.15). The unit exergy<br />

consumption increase was obvious ( ∆k<br />

= 0.200, see table A1.13). The intrinsic<br />

malfunction was therefore 409 kW, <strong>and</strong> the impact on fuel associated with the<br />

inefficiency is 608 kW (the total impact on fuel taking for the whole system is<br />

750 kW).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

345


TABLE A1.9 F-P design values.<br />

346<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.10 F-P values with inefficiency in FP: -12% in its efficiciency.<br />

Effect <strong>of</strong> feed pump isoentropic efficiency<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

347


TABLE A1.11 KP matrix in design (MCR case).<br />

348<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.12 KP matrix when the inefficiency in FP is detected.<br />

Effect <strong>of</strong> feed pump isoentropic efficiency<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

349


TABLE A1.13 Variation <strong>of</strong> the KP matrix when the FP is working improperly.<br />

350<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.14 Irreversibility matrix with -12% in the FP efficiency.<br />

Effect <strong>of</strong> feed pump isoentropic efficiency<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

351


TABLE A1.15 Dysfunction table <strong>and</strong> malfunction array when the FP is working with 12% lower efficiency.<br />

352<br />

<strong>Thermoeconomic</strong> diagnosis<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.16 Malfunction matrix when the efficiency <strong>of</strong> the FP varies 1%.<br />

Effect <strong>of</strong> feed pump isoentropic efficiency<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

353


FIGURE A1.6 Impact on fuel <strong>analysis</strong> when a inefficiency in FP is detected.<br />

354<br />

<strong>Thermoeconomic</strong> diagnosis<br />

FIGURE A1.7 Irreversibility <strong>analysis</strong> with the irreversibility in FP.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Effect <strong>of</strong> feed pump isoentropic efficiency<br />

The malfunction induced in the boiler was mainly due to the large amount <strong>of</strong><br />

product it generates (210 MW), since its efficiency <strong>and</strong> unit exergy consumption are<br />

not varied (the ∆k component is close to zero, see the ∆ 〈KP〉 matrix in table A1.13).<br />

This provokes a malfunction <strong>of</strong> –80 kW <strong>and</strong> an impact on fuel <strong>of</strong> only –69 kW.<br />

The irreversibility <strong>analysis</strong> shows that the dysfunctions associated with the boiler<br />

<strong>and</strong> condenser were the highest (754 <strong>and</strong> –472 kW respectively) <strong>and</strong> were generated<br />

by the feed pump. The weight coefficients φ ij (see table A1.14) were quite high in the<br />

rows corresponding to boiler <strong>and</strong> condenser (the unit consumption was changed in<br />

this case because the final products <strong>of</strong> these components had to increase 370 <strong>and</strong><br />

550 kW respectively to maintain the net output power). In these rows, the pump<br />

inefficiency dysfunction was provoked by varying the unit exergy consumption <strong>of</strong><br />

components more related to other components (i.e., the boiler <strong>and</strong> the condenser).<br />

Since the feed pump does not induce malfunctions <strong>and</strong> the model reacts linearly to<br />

variations in pump efficiency, the malfunction matrix is an exact tool to quantify<br />

additional fuel consumption for this inefficiency. Figure A1.8 demonstrates this<br />

linear behavior at the extraction mode load (122 MW).<br />

FIGURE A1.8 Effect <strong>of</strong> feed pump efficiency on fuel consumption. Variational study in the MCR performance<br />

case.<br />

Inc. fuel consumption<br />

800<br />

kW<br />

600<br />

400<br />

200<br />

0<br />

-12 -10 -8 -6 -4 -2 0 2 4 6 8 10 12<br />

-200<br />

-400<br />

-600<br />

% eff. in FP<br />

The cost <strong>of</strong> electricity <strong>and</strong> water as a function <strong>of</strong> feed pump inefficiency was very<br />

clear <strong>and</strong> linear (see figures A1.9 <strong>and</strong> A1.10). We can save 0.000003 $/kW·h <strong>and</strong><br />

0.00006 $/m 3 in electricity <strong>and</strong> freshwater production with a 1% increase in pump<br />

efficiency. The relative effect on electricity (the effect per unit produced) is<br />

supposedly greater than the effect on water. For a constant yearly production, a 1%<br />

isoentropic efficiency implies a savings <strong>of</strong> 3,530 $/y in electricity <strong>and</strong> 1,260 $/y in<br />

water.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 355


<strong>Thermoeconomic</strong> diagnosis<br />

FIGURE A1.9 Effect <strong>of</strong> pump inefficiency on electricity cost (MCR performance case).<br />

Electricity cost<br />

0,03784<br />

0,03782<br />

0,03780<br />

0,03778<br />

0,03776<br />

% eff. in FP<br />

-12 -10 -8 -6 -4 -2 0 2 4 6 8 10 12<br />

FIGURE A1.10 Water cost when the efficiency <strong>of</strong> the feed pump is varied.<br />

The main results <strong>of</strong> the inefficiency <strong>analysis</strong> were:<br />

• As expected, the effect <strong>of</strong> the feed pump inefficiency was only local. The<br />

increase in feedwater temperature leaving the pump was almost insignificant.<br />

The additional electrical consumption <strong>of</strong> the pump did not change the steam<br />

cycle behavior. The additional fuel supplied the extra electrical consumption <strong>of</strong><br />

the feed pump. The effect <strong>of</strong> this inefficiency is not as important as inefficiencies<br />

in other components, such as the steam turbine sections (less than 5,000 $/y in<br />

the <strong>combined</strong> production <strong>of</strong> water <strong>and</strong> electricity).<br />

• The feed pump is not strategic in a power plant. Its effects need only be<br />

considered if an inefficiency stops the plant because <strong>of</strong> a broken pump<br />

component (i.e. the linearity <strong>of</strong> the variational <strong>analysis</strong> is not valid).<br />

• The product <strong>of</strong> the steam power plant must be the net output power. The effect <strong>of</strong><br />

this inefficiency is not clearly noted in the system if gross output power is<br />

maintained.<br />

356 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

$/kWh<br />

Water cost<br />

1,2730<br />

1,2725<br />

1,2720<br />

1,2715<br />

1,2710<br />

$/m3<br />

% eff. in FP<br />

-12 -10 -8 -6 -4 -2 0 2 4 6 8 10 12


Effect <strong>of</strong> an inefficiency in the first section <strong>of</strong> the high-pressure turbine (HPT1)<br />

A1.3 Effect <strong>of</strong> an inefficiency in the first section <strong>of</strong> the<br />

high-pressure turbine (HPT1)<br />

The physical effects <strong>of</strong> an inefficiency in a turbine section are described in section<br />

7.3.2.1 as intrinsic malfunctions (steam path degradation, etc). HPT1 contains the<br />

governing section which is also affected by the control valves. Although the steam<br />

power plant always works at constant pressure, an intrinsic malfunction in the Curtis<br />

blade or wheels induces malfunctions downstream (see section 7.3.2.1). This varies<br />

steam conditions downstream because the steam conditions exiting HPT1 are<br />

changed, even though the HPT exhaust pressure remains constant. As this steam<br />

passes through the rest <strong>of</strong> turbine sections, they should also be affected, although<br />

their isoentropic efficiencies remain almost constant due to a constant pressure ratio.<br />

Conditions <strong>of</strong> low-pressure steam are slightly varied as the HPT exhaust values are<br />

sent to the MSF unit. The exhaust pressure remains constant by definition. The<br />

system can only respond to the inefficiency by producing additional live steam to<br />

maintain output power. This extra steam is proportionally spread over the steam<br />

cycle so no new induced malfunctions (in pre-heaters or pumps) arise. The noninefficient<br />

turbine sections produce the power that the inefficient section cannot<br />

produce.<br />

HPT1 efficiency was varied to observe its effect on other plant components <strong>and</strong><br />

additional consumption. We considered a production <strong>of</strong> 122 MW in extraction mode<br />

with a 5% decrease in isoentropic efficiency. The diagnosis was also developed at a<br />

similar degree <strong>of</strong> inefficiency for 60 MW (parallel mode), 90 MW (extraction mode)<br />

<strong>and</strong> 140 MW (condensing mode).<br />

Tables A1.17-A1.24 show, step by step, the methodology applied in the previous<br />

sections. Tables A1.17 <strong>and</strong> A1.18 are the F-P definition tables <strong>of</strong> the design <strong>and</strong><br />

inefficient situation, tables A1.19 <strong>and</strong> A1.20 are the 〈KP〉 matrices. The ∆ 〈KP〉<br />

matrix <strong>and</strong> the irreversibility matrix are depicted in tables A1.21 <strong>and</strong> A1.22, <strong>and</strong> the<br />

[DF] matrix <strong>and</strong> the malfunction matrix are shown in tables A1.23 <strong>and</strong> A1.24.<br />

Figures A1.11 <strong>and</strong> A1.12 show the impact on fuel <strong>and</strong> irreversibility increase<br />

<strong>analysis</strong> <strong>of</strong> the inefficiency.<br />

An inefficiency in an component producing an important part <strong>of</strong> the final product<br />

should have important consequences. Other components have to readapt the turbine<br />

section to maintain electricity production <strong>and</strong> improve their efficiency (turbine<br />

sections) or consume more resources (boiler). A inefficiency diagnosis will explain<br />

these ideas.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 357


TABLE A1.17 F-P values without any inefficiency. MCR case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

358 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.18 F-P values when the HPT1 decreases 5% its efficiency (MCR case).<br />

Effect <strong>of</strong> an inefficiency in the first section <strong>of</strong> the high-pressure turbine (HPT1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 359


TABLE A1.19 KP matrix in design (MCR case).<br />

<strong>Thermoeconomic</strong> diagnosis<br />

360 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.20 KP matrix when the inefficiency in HPT1 is 5% in its efficiency.<br />

Effect <strong>of</strong> an inefficiency in the first section <strong>of</strong> the high-pressure turbine (HPT1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 361


TABLE A1.21 Variation <strong>of</strong> the KP with the inefficiency in HPT1 (MCR case).<br />

<strong>Thermoeconomic</strong> diagnosis<br />

362 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.22 Irreversibility matrix with the inefficiency in HPT1 (MCR case).<br />

Effect <strong>of</strong> an inefficiency in the first section <strong>of</strong> the high-pressure turbine (HPT1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 363


TABLE A1.23 Dysfunction/malfunction table when the efficiency <strong>of</strong> the HPT1 is decreased 5%.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

364 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.24 Malfunction matrix when the efficiency <strong>of</strong> the HPT1 is varied 1%.<br />

Effect <strong>of</strong> an inefficiency in the first section <strong>of</strong> the high-pressure turbine (HPT1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 365


FIGURE A1.11 Impact on fuel <strong>analysis</strong> when the HPT1efficiency is 5% less than the expected.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

366 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

FIGURE A1.12 Irreversibility <strong>analysis</strong> with the inefficiency in HPT1.


Effect <strong>of</strong> an inefficiency in the first section <strong>of</strong> the high-pressure turbine (HPT1)<br />

The malfunctions <strong>of</strong> this inefficiency will be analyzed using table A1.23. The<br />

components with a malfunction that surpasses a non-negligible quantity are the<br />

inefficient component (HPT1), the boiler <strong>and</strong> the MSF unit. First we will explain the<br />

inefficient component. If the isoentropic efficiency <strong>of</strong> a turbine section decreases,<br />

the expansion line is moved away from the reversible process. The irreversibility in<br />

the section increases by 1,667 kW. Since the turbine exhaust has a higher enthalpy<br />

(see the h-s diagram), the output power strongly decreases with respect to the design<br />

situation (2,220 kW). This means that unit exergy consumption increases <strong>and</strong> the<br />

product decreases. The ∆ 〈KP〉 component <strong>of</strong> HPT1 was ∆k = 0.039 (see table<br />

A1.21). The intrinsic malfunction was 1,948 kW, <strong>and</strong> the impact on fuel due to the<br />

inefficiency was 2,825 kW. Since the total impact on fuel in the plant was 3,732 kW,<br />

this parameter could be considered local to the system.<br />

However, the malfunction associated with the MSF unit is negative<br />

(MF = --280 kW), if we assume that the water produced <strong>and</strong> the condensate returned<br />

to the deaerator are constant. The is because the end point <strong>of</strong> the expansion line is<br />

located in HPT. Steam leaving HPT has a higher enthalpy but also a higher entropy.<br />

The energy needed by the MSF unit also increases, decreasing efficiency. The<br />

generated negentropy in the MSF unit is considered a secondary product <strong>of</strong> the<br />

component <strong>and</strong> is beneficial (see section 7.3.2.1), so the final variation <strong>of</strong> the unit<br />

exergy consumption is negative (∆k = –0.041).<br />

The malfunction associated with the boiler is also negative. As its product exergy<br />

flow is huge, the induced malfunction is –242 kW, although its unit exergy<br />

consumption did not change very much (∆k = –0.0011, see table A1.21). The reason<br />

is the increased feedwater temperature entering the boiler due to the additional<br />

steam required by the steam power plant to maintain the electricity production in the<br />

inefficient situation. The additional fuel consumed is not used for the same<br />

temperature rise in the boiler with respect to the design conditions. The impact on<br />

fuel associated with the boiler is –175 kW, but the irreversibility in the component<br />

increases 3,341 kW. The last assumption is a consequence <strong>of</strong> the dysfunction<br />

<strong>analysis</strong> explained below.<br />

The dysfunction <strong>analysis</strong> is quite similar to when other components suffer<br />

inefficiencies. Once again, the two components suffering from the dysfunctions<br />

generated by the components with an inefficiency are the boiler <strong>and</strong> the condenser.<br />

In both components the highest dysfunction is provoked by the inefficient<br />

component (2,616 kW for the boiler <strong>and</strong> –1,795 kW for the condenser), that is, the<br />

component with the intrinsic <strong>and</strong> greatest malfunction. The sum <strong>of</strong> the dysfunctions<br />

generated by the other components is the irreversibility increase associated with<br />

each component. For example, the total dysfunction generated in the boiler is<br />

3,583 kW <strong>and</strong> its production is increased by 1,790 kW to maintain the final<br />

production <strong>of</strong> the power plant.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 367


<strong>Thermoeconomic</strong> diagnosis<br />

Figure A1.13 shows the linearity <strong>of</strong> the model when the isoentropic efficiency is<br />

varied from –5 to +5 %, in extraction mode under MCR (122 MW), in this figure the<br />

total impact on fuel associated to this inefficiency is analyzed.<br />

FIGURE A1.13 Model linearity with respect to an inefficiency in HPT1.<br />

Inc. fuel consumption<br />

4000<br />

kW<br />

3000<br />

2000<br />

1000<br />

-5 -4 -3 -2 -1 0 1 2 3 4 5<br />

-1000<br />

-2000<br />

-3000<br />

-4000<br />

0<br />

% eff. in HPT1<br />

The model behaved more <strong>of</strong> less linearly to variations <strong>of</strong> the inefficiency using the<br />

simulator. Thus, the malfunction matrix can be used to predict the impact on fuel.<br />

Since the inefficiency does not provoke any important induced malfunctions in other<br />

components, the malfunction matrix could also be used when several inefficiencies<br />

are occurring in different components.<br />

FIGURE A1.14 Cost <strong>of</strong> electricity depending on the degree <strong>of</strong> inefficiency applied to HPT1 (MCR case).<br />

% eff. in HPT1<br />

Electricity cost<br />

0,0381<br />

$/kWh<br />

0,0380<br />

0,0379<br />

0,0378<br />

0,0377<br />

0,0376<br />

0,0375<br />

-5 -4 -3 -2 -1 0 1 2 3 4 5<br />

The cost <strong>of</strong> electricity <strong>and</strong> water as a function <strong>of</strong> the isoentropic efficiency <strong>of</strong> HPT1<br />

illustrates its effect (see figures A1.14 <strong>and</strong> A1.15). A 1% decrease in the isoentropic<br />

efficiency in HTP1 means an additional cost <strong>of</strong> 0.00004 $/kW·h (44,900 $/y) in<br />

electricity <strong>and</strong> 0.0005 $/m 3 in water (11,800 $/y). Clearly the inefficiency should be<br />

368 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Effect <strong>of</strong> inefficiency in the first section <strong>of</strong> the low-pressure turbine (LPT1)<br />

corrected to avoid additional costs, because the first section is responsible for a high<br />

percentage <strong>of</strong> the total electricity produced in the steam power plant.<br />

FIGURE A1.15 Cost <strong>of</strong> water when the isoentropic efficiency is varied from –5% to 5% with respect to design<br />

efficiency (MCR case).<br />

% eff. in HPT1<br />

Water cost<br />

1,275<br />

$/m3<br />

1,274<br />

1,273<br />

1,272<br />

1,271<br />

1,270<br />

1,269<br />

1,268<br />

-5 -4 -3 -2 -1 0 1 2 3 4 5<br />

The most important results derived from the <strong>analysis</strong> <strong>of</strong> this inefficiency include:<br />

• HPT1 is very important in terms <strong>of</strong> additional fuel consumption <strong>and</strong> cost <strong>of</strong><br />

water <strong>and</strong> electricity (more than 55,000 $/y savings in the two products when the<br />

inefficiency is improved by only 1%).<br />

• The steam conditions exiting HPT1 also affect (to a lesser degree) some other<br />

components receiving that steam, i.e. the MSF unit. In any case, the inefficiency<br />

could be considered local to the turbine section.<br />

• The HPT1 inefficiency should be avoided, even if the turbine needs repair to<br />

prevent against inefficiencies or failures, since the savings would be quickly<br />

recovered.<br />

A1.4 Effect <strong>of</strong> inefficiency in the first section <strong>of</strong> the lowpressure<br />

turbine (LPT1)<br />

The low-pressure turbine has only two sections in the power plant configuration.<br />

Unless the plant is working at condensing mode, the amount <strong>of</strong> steam sent to this<br />

turbine is very low. Thus, an inefficiency in this section should have less effect than<br />

other inefficiencies in the turbine sections. The induced malfunctions should be<br />

detected in the second section <strong>of</strong> the low-pressure turbine. The degradation process<br />

could be accelerated if the last section <strong>of</strong> the low-pressure turbine has to work as a<br />

compressor when the amount <strong>of</strong> steam diverted to this section is so low that the<br />

steam cannot overcome the mechanical losses <strong>of</strong> the turbine. But this section also<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 369


<strong>Thermoeconomic</strong> diagnosis<br />

suffers from induced malfunctions form HPT (according to the definition <strong>of</strong> induced<br />

<strong>and</strong> intrinsic malfunctions by Royo (1994) for a steam turbine). The amount <strong>of</strong><br />

steam to the MSF unit gives the pressure <strong>of</strong> the steam leaving the high-pressure<br />

turbine. Some part <strong>of</strong> this steam is also introduced in the low-pressure turbine.<br />

Finally, the atmospheric conditions control the exhaust pressure <strong>of</strong> the turbine<br />

making the behavior <strong>of</strong> this section strongly dependent the ambient temperature.<br />

This inefficiency <strong>analysis</strong> was performed for the MCR case (122 MW power<br />

production with an extraction to the MSF unit <strong>of</strong> 89.68 kg/s). The physical effects <strong>of</strong><br />

these inefficiencies were translated into malfunctions <strong>and</strong> additional fuel<br />

consumption. The isoentropic efficiency <strong>of</strong> this section was 15% lower than the<br />

design efficiency (about the 76%).<br />

Tables A1.25 <strong>and</strong> A1.26 show the F-P values <strong>of</strong> the <strong>simulation</strong> corresponding to the<br />

design <strong>and</strong> inefficient cases. If we apply the <strong>analysis</strong> for other operating modes<br />

(condensing or parallel mode, or 140 <strong>and</strong> 60 MW <strong>of</strong> output power, respectively), the<br />

productive structure changes (see section 7.1), <strong>and</strong> the F-P definitions <strong>and</strong> the rest <strong>of</strong><br />

matrices are different than in these examples. Tables A1.27 <strong>and</strong> A1.28 include the<br />

〈KP〉 matrices dividing the fuels <strong>and</strong> products <strong>of</strong> each component. Table A1.29 is the<br />

∆ 〈KP〉 matrix composed by the subtraction <strong>of</strong> the two last matrices, <strong>and</strong> table A1.30<br />

is the irreversibility matrix |I〉. Finally, table A1.31 is the dysfunction/malfunction<br />

table, <strong>and</strong> table A1.32 is the malfunction matrix associated with the inefficiency in<br />

LPT1. Figures A1.16 <strong>and</strong> A1.17 include the impact on fuel <strong>and</strong> the increase <strong>of</strong><br />

irreversibility.<br />

An inefficiency in LPT1 is less important than in HPT1 in a co-generation plant. The<br />

HPT does not detect an inefficiency. The conditions <strong>of</strong> the steam downstream the<br />

inefficient component do vary but the exhaust pressure is controlled by the external<br />

temperature <strong>and</strong> does not vary, although the exhausted vapor to the condenser can<br />

vary its humidity. Some other turbine sections have to readapt their production to<br />

produce the electricity required, as their efficiencies do not vary when some amount<br />

<strong>of</strong> extra live steam is dem<strong>and</strong>ed to the boiler.<br />

In the malfunction array <strong>of</strong> this inefficiency, the inefficient component (LPT1) <strong>and</strong><br />

the first section <strong>of</strong> the high-pressure turbine have a higher malfunction than the<br />

minimum accuracy <strong>of</strong> the simulator. The physical interpretation <strong>of</strong> these<br />

malfunctions will be connected. The irreversibility <strong>of</strong> the steam expansion increases<br />

in the inefficient component <strong>of</strong> LTP1 (∆I = 2,062 kW, see table A1.31), when the<br />

isoentropic efficiency decreases. The electricity production <strong>of</strong> the component<br />

reduces by 1,230 kW, <strong>and</strong> its variation <strong>of</strong> unit exergy consumption is ∆k = 0.522<br />

(see table A1.29). The last assumptions result in an intrinsic malfunction (that is, the<br />

malfunction created in the inefficient component <strong>of</strong> a system) <strong>of</strong> 3,408 kW. The total<br />

malfunction associated with the whole plant is 3,260 kW. Clearly this inefficiency<br />

does not provoke any induced malfunctions in the rest <strong>of</strong> the plant components.<br />

370 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.25 F-P values in design (MCR case).<br />

Effect <strong>of</strong> inefficiency in the first section <strong>of</strong> the low-pressure turbine (LPT1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 371


TABLE A1.26 F-P values with the inefficiency in LPT1, MCR case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

372 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.27 KP matrix in design, MCR case.<br />

Effect <strong>of</strong> inefficiency in the first section <strong>of</strong> the low-pressure turbine (LPT1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 373


TABLE A1.28 KP matrix when the efficiency in the LPT1 is decreased 15%, MCR case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

374 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.29 Variation <strong>of</strong> the KP matrix with an inefficiency in LPT1, MCR case.<br />

Effect <strong>of</strong> inefficiency in the first section <strong>of</strong> the low-pressure turbine (LPT1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 375


TABLE A1.30 Irreversibility matrix with the efficiency <strong>of</strong> the LPT1 decreased 15%, MCR case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

376 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.31 Dysfunction/malfunction table for an inefficiency in the LPT1 (15%), MCR case.<br />

Effect <strong>of</strong> inefficiency in the first section <strong>of</strong> the low-pressure turbine (LPT1)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 377


TABLE A1.32 Malfunction matrix when the efficiency <strong>of</strong> the LPT1 is varied 1%, MCR case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

378 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE A1.16 Impact on fuel <strong>analysis</strong>, section A1.4.<br />

Effect <strong>of</strong> inefficiency in the first section <strong>of</strong> the low-pressure turbine (LPT1)<br />

FIGURE A1.17 Irreversibility <strong>analysis</strong> in section A1.4.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 379


<strong>Thermoeconomic</strong> diagnosis<br />

HPT1 has a negative malfunction <strong>of</strong> 215 kW <strong>and</strong> ∆k = –0.004 (see table A1.29).<br />

This negative value is explained in the mathematical model <strong>of</strong> the steam turbine. The<br />

amount <strong>of</strong> steam entering the Curtis blade is higher than expected <strong>and</strong> the section<br />

operates more efficiently when the steam leaving this section is slightly increased.<br />

The total impact on fuel associated with this effect was –281 kW.<br />

The dysfunction <strong>analysis</strong> applied to this inefficiency is very illustrative. Only the<br />

boiler <strong>and</strong> condenser suffer dysfunctions generated by the components with<br />

malfunctions: HPT1 <strong>and</strong> LTP1. In both cases these components have to readapt<br />

production by 1,470 <strong>and</strong> 2,430 kW respectively, to maintain the additional<br />

production required by the first section <strong>of</strong> the low-pressure turbine. Since these two<br />

components redistribute their products over the rest <strong>of</strong> the components, their φ ij<br />

coefficients are not zero. If there is a ∆k ij coefficient whose value is not zero, the<br />

dysfunction generated by the last component in the first two components is<br />

significant. The rest <strong>of</strong> components do not have any important dysfunction worth<br />

mentioning in our <strong>analysis</strong>.<br />

Figure A1.18 shows the effect <strong>of</strong> varying the efficiency in this turbine section around<br />

the design point. The efficiency was varied from –15 to +15% with respect to this<br />

point. Since the model was linear with respect to the inefficiency, the malfunction<br />

matrix (table A1.32) can be used to quantify the additional fuel consumption by<br />

multiplying this matrix by the product <strong>and</strong> the unit exergy cost <strong>of</strong> every component.<br />

With this inefficiency there were no induced malfunctions (isolated component),<br />

making the malfunction matrix an exact guide to predict the increment on fuel<br />

consumption.<br />

FIGURE A1.18 Effect on the fuel consumption when the degree <strong>of</strong> inefficiency in the LPT1 is varied from the<br />

design point (MCR case).<br />

Inc. fuel consumption<br />

4000<br />

2000<br />

-2000<br />

-4000<br />

0<br />

% eff. in LPT1<br />

-15 -12 -9 -6 -3 0 3 6 9 12 15<br />

The monetary cost (including the capital cost <strong>and</strong> device maintenance) <strong>of</strong> water <strong>and</strong><br />

electricity is one <strong>of</strong> the consequences <strong>of</strong> the diagnosis <strong>of</strong> the plant with respect to an<br />

component inefficiency. Figures A1.19 <strong>and</strong> A1.20 show how the cost in electricity<br />

380 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

kW


Effect <strong>of</strong> inefficiency in the first section <strong>of</strong> the low-pressure turbine (LPT1)<br />

increases 0.000015 $/kWh <strong>and</strong> the water increases 0.00006 $/m 3 when the LTP1<br />

isoentropic efficiency decreases by 1%. In a year, at 122 MW <strong>and</strong> 2,400 T/h,<br />

15,000 $ <strong>and</strong> 1,280 $ are saved in electricity <strong>and</strong> water costs.<br />

FIGURE A1.19 Cost <strong>of</strong> electricity for inefficiencies in LPT1 (MCR case).<br />

Electricity cost<br />

0,0381<br />

$/kWh<br />

0,0379<br />

0,0377<br />

0,0375<br />

-15 -12 -9 -6 -3 0 3 6 9 12 15<br />

FIGURE A1.20 Water cost per cubic meter for inefficiencies in LPT1. 122 MW in extraction mode (MCR case).<br />

% eff. in LPT1<br />

Water cost<br />

1,2730<br />

$/m3<br />

1,2725<br />

1,2720<br />

1,2715<br />

1,2710<br />

1,2705<br />

This section demonstrated that:<br />

% eff. in LPT1<br />

-15 -12 -9 -6 -3 0 3 6 9 12 15<br />

• The behavior <strong>of</strong> LPT1 is linear when its efficiency is varied within allowable<br />

limits. It does not induce any significant malfunctions in other plant components,<br />

following the trend in other examples.<br />

• As predicted in the first paragraph <strong>of</strong> this section, the cost <strong>of</strong> water <strong>and</strong><br />

electricity were not affected as much as by inefficiencies in HPT (only 16,300 $/<br />

y are saved in both products if the isoentropic efficiency is improved 1%).<br />

Therefore, the effect <strong>of</strong> an inefficiency in the turbine is proportional to the<br />

amount <strong>of</strong> steam entering the turbine section.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 381


<strong>Thermoeconomic</strong> diagnosis<br />

• The most dangerous problem associated with inefficiencies in LPT is the steam<br />

quality when the efficiency is increased. Low quality steam can damage the<br />

wheels <strong>of</strong> the condensing turbine. The variational <strong>analysis</strong> can also be broken<br />

when the inefficiency provokes a non-linear system response.<br />

A1.5 Effect <strong>of</strong> the cleaning ball system in the recovery<br />

section<br />

The recovery section is the most important component <strong>of</strong> the MSF unit. Therefore,<br />

the cleaning ball system inside the distiller tubes could provoke several malfunctions<br />

in other plant components. In this section we analyze the fouling reduction effect.<br />

The benefits <strong>of</strong> reducing fouling in the reject section can be translated into the<br />

physical response <strong>of</strong> the MSF unit. First, an <strong>analysis</strong> was done keeping the control<br />

parameters constant (SR, R, F). If the fouling is decreased in the recovery section,<br />

heat transfer inside the tubes is increased <strong>and</strong> the inter-stage temperature difference<br />

between the vapor <strong>and</strong> cooling brine decreases. This raises the temperature <strong>of</strong><br />

cooling brine <strong>and</strong> decreases the flashing brine <strong>and</strong> released vapor. But the cooling<br />

brine goes to the brine heater since it is hotter than in design. Finally, the cooling<br />

brine flow enters the recovery section at a higher temperature than expected. In the<br />

final stages <strong>of</strong> the recovery section, both distillate <strong>and</strong> flashing temperatures are<br />

reduced by the effect <strong>of</strong> the fouling inside the recovery tubes. The flash range <strong>of</strong> the<br />

distillers is increased in the two limits <strong>and</strong> the distillate produced in the MSF unit is<br />

higher than in design. The control parameters <strong>of</strong> the MSF unit (seawater to reject<br />

SR, recycle brine R or make-up feed F) must be reduced if the distillate product is to<br />

be maintained (although the distillate temperature leaving the unit could be reduced)<br />

<strong>and</strong>, indirectly, the amount <strong>of</strong> steam consumed in the heater. The diagnosis<br />

mathematically explains the physical effects.<br />

Tables A1.33 <strong>and</strong> A1.34 show the F-P definition matrices following the productive<br />

structure in section 7.1. Then, the 〈KP〉 matrices from the last two matrices (tables<br />

A1.35 <strong>and</strong> A1.36) are shown. The ∆ 〈KP〉 matrix (table A1.37) is obtained by<br />

subtracting tables A1.35 <strong>and</strong> A1.36. The irreversibility matrix |I〉 (table A1.38) <strong>and</strong><br />

the malfunction/dysfunction table is shown in table A1.39. The example analyzed<br />

produced 1.900 T/h with 32 ºC seawater (nominal-temperature operation <strong>of</strong> the MSF<br />

distillers in summer). Figures A1.21 <strong>and</strong> A1.22 show the impact on fuel <strong>analysis</strong> <strong>and</strong><br />

the increase <strong>of</strong> the irreversibility in the MSF components.<br />

382 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.33 F-P values in design, NTOS case.<br />

Effect <strong>of</strong> the cleaning ball system in the recovery section<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 383


TABLE A1.34 F-P values with fouling in RCS=0, NTOS case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

384 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.35 KP matrix in design, NTOS case.<br />

Effect <strong>of</strong> the cleaning ball system in the recovery section<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 385


TABLE A1.36 KP matrix with an inefficiency in RCS, NTOS case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

386 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


FIGURE A1.21 Impact on fuel <strong>analysis</strong> in section A1.5.<br />

Effect <strong>of</strong> the cleaning ball system in the recovery section<br />

FIGURE A1.22 Irreversibility increase in section A1.5.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 387


<strong>Thermoeconomic</strong> diagnosis<br />

In the malfunction <strong>analysis</strong>, only the inefficient component has an intrinsic<br />

malfunction <strong>of</strong> –1,570 kW. This low value can be explained physically. The fouling<br />

reduction inside the recovery tubes improves the heat transfer coefficient in those<br />

stages, reducing the thermal irreversibility (∆I = –2,132 kW, see table A1.39) <strong>and</strong><br />

also the flows recirculating in the recovery section in order to maintain final<br />

production. This means that the variation <strong>of</strong> the unit exergy consumption is<br />

∆k = --0.194, see table A1.37. The impact on fuel associated with the inefficient<br />

component is –4,279 kW.<br />

The brine heater has an induced malfunction <strong>of</strong> –626 kW. The brine entering the<br />

heater has a higher temperature due to improved heat transmission in the recovery<br />

section, but the temperature entering the distiller is reduced by 0.3 ºC. The brine<br />

heater needs less steam to heat the cooling brine, considering that the recycle brine<br />

flow is also reduced to maintain distillate production. This means that the<br />

irreversibility generated in the heater is also reduced by ∆I = –1,131 kW, <strong>and</strong><br />

therefore the variation <strong>of</strong> the unit exergy consumption (the ∆ 〈KP〉 coefficient is<br />

∆k = –0.0149, see table A1.37).<br />

The inefficiency in the recovery induces a –540 kW malfunction in the reject<br />

section. The distillate flow leaving the section depends on the temperature <strong>of</strong> the<br />

flashing brine <strong>and</strong> distillate entering the plant (both temperatures decrease 2.6 ºC)<br />

<strong>and</strong> the recycle brine to the distiller (which is reduced 263 T/h). The energy required<br />

to produce the distillate is lower than the design value <strong>and</strong> the irreversibility<br />

generated in this section (∆I = –554 kW, see table A1.39). The unit exergy<br />

consumption <strong>of</strong> the reject is reduced because the amount <strong>of</strong> resources to distillate<br />

the freshwater is lower (∆k = –0.079, see table A1.37). As the distillate is produced<br />

at a considerable exergy cost (see the last row <strong>of</strong> table A1.38 for the exergy cost <strong>of</strong><br />

each component), 5,408 kW <strong>of</strong> fuel was saved with this induced malfunction.<br />

The component suffering the highest malfunction is the fictitious device (FD),<br />

included in the productive structure to quantify the flows sent to sea: blowdown <strong>and</strong><br />

reject cooling seawater. The malfunction associated with this component (7,850 kW,<br />

see table A1.39) can only be explained by the thermoeconomic model. Its product<br />

(the fuel consumed in the MSF unit to produce freshwater, i.e. the steam coming<br />

from the steam power plant) is obviously decreased with the use <strong>of</strong> the cleaning ball<br />

system. The exergy flow <strong>of</strong> the steam to the MSF unit decreases 9,250 kW. The<br />

increase in unit exergy consumption is ∆k = 0.170 <strong>and</strong> the impact on fuel associated<br />

with this component is 13,249 kW. It is clearly not convenient to use non-physical<br />

components in the productive structure <strong>of</strong> the system because their associated<br />

malfunctions <strong>and</strong> dysfunctions are quite difficult to explain physically.<br />

388 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.37 Variation <strong>of</strong> the KP matrix when the fouling in RCS is neglected.<br />

Effect <strong>of</strong> the cleaning ball system in the recovery section<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 389


TABLE A1.38 Irreversibility matrix without fouling in RCS.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

390 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.39 Dysfunction/malfunction table without fouling in RCS, NTOS case.<br />

Effect <strong>of</strong> the cleaning ball system in the recovery section<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 391


<strong>Thermoeconomic</strong> diagnosis<br />

The mixer is also a non-physical device in the last stage <strong>of</strong> the reject section. It<br />

models the mixing process between the make-up feed <strong>and</strong> the brine flashing in the<br />

last stage <strong>of</strong> the reject section. It has a negative induced malfunction <strong>of</strong> 765 kW due<br />

to the reduction <strong>of</strong> irreversibility generated (∆I = –735 kW, see table A1.39) in the<br />

mixing process (the two mixed flows are reduced in quantity <strong>and</strong> energy). The<br />

efficiency <strong>of</strong> the process is therefore improved, with a unit exergy consumption<br />

variation <strong>of</strong> ∆k = –0.013 (see table A1.37). The impact on fuel associated with the<br />

‘benefunction’ in the mixer is –996 kW.<br />

Now the dysfunction <strong>analysis</strong> will be introduced. Components suffering a important<br />

malfunction clearly induce a large dysfunction in the rest <strong>of</strong> components. For<br />

example, the main dysfunctions in the fictitious device are generated by itself<br />

(5,016 kW), the heater (–1,226 kW), the recovery section (–2,534 kW), the mixer<br />

(--212 kW) <strong>and</strong> the reject section (–3,022 kW). The value <strong>of</strong> the dysfunction is<br />

proportional to the malfunction in each component. The dysfunction in a component<br />

due to the junctions <strong>of</strong> the productive structure must be distributed to the<br />

components supplying the junction. The total dysfunctions generated by each<br />

component were 5,398 kW for the FD, –1,263 kW for the heater, –2,708 kW for the<br />

recovery section, –230 kW for the mixer <strong>and</strong> finally –4,867 kW for the reject<br />

section. The temperature pr<strong>of</strong>ile change in the distillers provokes differences in the<br />

exergy <strong>of</strong> products leaving each component to readapt the final production <strong>of</strong><br />

distilled water.<br />

The previous <strong>analysis</strong> kept the final product <strong>of</strong> the system constant (distillate water).<br />

As mentioned in previous sections, the simulator can maintain the mass flow rate in<br />

the distiller but it cannot maintain the exergy <strong>of</strong> this flow. As in this case, the<br />

temperature <strong>of</strong> distillate leaving the MSF unit is reduced by 1.3 ºC. The impact on<br />

fuel associated with the variation <strong>of</strong> the final product is an astonishing –4,337 kW!<br />

This value is similar to the total impact on fuel associated with the unit exergy<br />

consumption variation inside the MSF unit (–5,336 kW).<br />

The variational <strong>analysis</strong> <strong>of</strong> this inefficiency involves the linear behavior <strong>of</strong> the<br />

model, as in the figure A1.23, where the total impact on fuel saved (including the<br />

final product variation) with decreased fouling is depicted from the design value to<br />

total absence. In this section we analyzed nominal production (1,900 T/h) with 32 ºC<br />

seawater <strong>and</strong> the typical exergy costs <strong>of</strong> electricity <strong>and</strong> steam obtained in the power<br />

plant <strong>analysis</strong>.<br />

The inefficiency diagnosis can also be quantified in monetary terms. The cost <strong>of</strong><br />

water depending on fouling helps plant managers develop a maintenance plan to<br />

operate under the best conditions. In this case (see figure A1.24) the cost <strong>of</strong> a cubic<br />

meter <strong>of</strong> water decreased 0.0069 $ when the fouling in this section decreased<br />

0.00001 m 2 ·K/W. This value is very high (115,000 $ a year) <strong>and</strong> implies that the<br />

cleaning ball system should always operate in the recovery section.<br />

392 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Effect <strong>of</strong> the cleaning ball system in the recovery section<br />

FIGURE A1.23 Effect on fuel consumption when the fouling in recovery section is gradually decreased. 1,900 T/<br />

h <strong>and</strong> 32º C seawater.<br />

0<br />

fouling*10-5 in RC<br />

0 3 6 9 12 15<br />

-4000<br />

-8000<br />

-12000<br />

-16000<br />

-20000<br />

-24000<br />

Inc. fuel consumption<br />

FIGURE A1.24 Cost <strong>of</strong> a cubic meter <strong>of</strong> water depending on the fouling in the recovery section.<br />

1,48<br />

1,45<br />

1,42<br />

1,39<br />

1,36<br />

$/m3<br />

kW<br />

Water cost<br />

fouling*10-5 in RC<br />

0 3 6 9 12 15<br />

The malfunction matrix (table A1.40) <strong>of</strong> the MSF unit with this inefficiency is a<br />

good tool to calculate the effect on natural gas. The malfunction matrix can be used<br />

because the model is linear with respect to the fouling in recovery. But the induced<br />

malfunctions produced by this inefficiency imply that the malfunction matrix can<br />

only be used for individual malfunctions.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 393


TABLE A1.40 Malfunction matrix when the fouling in RCS is varied 0.00001 m 2 K/W.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

394 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Effect <strong>of</strong> reject section fouling<br />

Summarizing the results:<br />

• The change <strong>of</strong> the temperature pr<strong>of</strong>ile by the fouling in the main flows <strong>of</strong> the<br />

MSF plant is responsible for the induced malfunctions in the distillers. Thus,<br />

each malfunction should be dealt individually. The values <strong>of</strong> the induced<br />

malfunctions surpass the intrinsic malfunction because the flows leaving <strong>and</strong><br />

entering the recovery section also pass through the reject or brine heater. The<br />

dysfunctions generated in the different components are also very important.<br />

• The increased heat transfer increases the production rate per stage in the distiller.<br />

This reduces the amount <strong>of</strong> resources to produce the same distillate. Since the<br />

cleaning ball system obviously saves fuel (115,000 $/y), it should operate<br />

continuously.<br />

• A large part <strong>of</strong> fuel saved with this inefficiency is due to the lower temperature<br />

<strong>of</strong> the distillate leaving the plant. But, in fact, the distillate temperature is now<br />

irrelevant (unless this energy is used by another process). So, this effect should<br />

not be considered during the <strong>analysis</strong>, although that temperature has a direct<br />

relationship with the other distiller temperatures.<br />

A1.6 Effect <strong>of</strong> reject section fouling<br />

Usually the cleaning ball system is not installed in the reject section since its<br />

seawater operating temperatures do not produce any scaling problems. But the<br />

biological activity <strong>of</strong> seawater intake can lead to dangerous bio-fouling in this<br />

section. The effect <strong>of</strong> installing a cleaning ball system here is similar to the recovery<br />

section. It reduces the interstage difference because the distillate temperature<br />

decreases <strong>and</strong> the cooling brine is heated to a higher temperature. Since the seawater<br />

temperature is imposed by the environment, the distillate temperature is forced to<br />

decrease when the heat transfer coefficient <strong>of</strong> each stage is increased, because the<br />

fouling inside the tubes is neglected. In this case, the flash range <strong>of</strong> the plant ∆T is<br />

higher because the lower limit <strong>of</strong> this range is decreased. A higher flash range<br />

implies a higher distillation per stage. If the control parameters <strong>of</strong> the plant are<br />

maintained, it can only produce additional freshwater with the help <strong>of</strong> the cleaning<br />

ball system. A lower recycle (R), seawater to reject (SR) <strong>and</strong> make-up feed (F) flow<br />

are needed to maintain the distillate production.<br />

As the brine heater is so far from the reject section, the temperature pr<strong>of</strong>iles <strong>of</strong> the<br />

cooling brine entering <strong>and</strong> leaving the heater do not vary considerably. The<br />

performance indexes or steam consumption <strong>of</strong> the plant are not expected to greatly<br />

improve.<br />

This system should not be used for several reasons based on thermoeconomic<br />

criteria. The example is the same as in previous sections: a water production <strong>of</strong><br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 395


<strong>Thermoeconomic</strong> diagnosis<br />

1.900 T/h with 32 ºC seawater <strong>and</strong> a fouling factor reduced to zero. Tables A1.41<br />

<strong>and</strong> A1.42 show the F-P definition applied to the design <strong>and</strong> inefficient case, tables<br />

A1.43 <strong>and</strong> A1.44 are the ∆ 〈KP〉 matrix made by using the previous tables, table<br />

A1.45 is the ∆ 〈KP〉 matrix <strong>and</strong> table A1.46 is the irreversibility matrix I containing<br />

the dysfunction coefficients <strong>and</strong> the exergy cost array. The dysfunction/malfunction<br />

matrix [DF]/MF is the table that resumes the final results <strong>of</strong> the thermoeconomic<br />

diagnosis applied to this inefficiency (table A1.47). The impact on fuel <strong>and</strong> the<br />

increase <strong>of</strong> irreversibility per component are shown in figures A1.25 <strong>and</strong> A1.26<br />

respectively.<br />

Although the reject section has three stages, the effect <strong>of</strong> fouling should be identical<br />

to the effect observed in the recovery section (17 stages). In this case, the<br />

temperature <strong>of</strong> cooling brine entering the distiller is given by the ambient<br />

conditions. The flashing <strong>and</strong> distillate temperatures would try to reach the cooling<br />

temperature flowing inside the tubes if the heat transfer were an ideal process. The<br />

symbolic formulation <strong>of</strong> thermoeconomics will give us the effects provoked by this<br />

inefficiency in the MSF unit.<br />

The most significant malfunctions are yet again located in the fictitious device,<br />

heater, recovery <strong>and</strong> reject sections <strong>and</strong> the mixer. The inefficient component<br />

<strong>analysis</strong> considers the cleaning ball system installed in the reject section.<br />

Suprisingly, the associated malfunction with no fouling in the reject is positive<br />

(49 kW). The ∆ 〈KP〉 component corresponding to its exergy unit consumption is<br />

∆k = 0.007 (see table A1.45). But this result is provoked by the assumptions adopted<br />

in the thermoeconomic model <strong>of</strong> the reject section. The part <strong>of</strong> the unit exergy<br />

consumption corresponding to the efficiency <strong>of</strong> the process (or the heat transfer<br />

improvement) is logically lower than the design situation (∆k 1 = –0.024). But the<br />

steam <strong>and</strong> brine needed for maintaining the vacuum inside the chambers is more or<br />

less independent from the distillate produced (i.e. is a constant value). As the<br />

product <strong>of</strong> the reject section decreases (the distillate temperature leaves the section<br />

at a lower temperature), the unit exergy consumption due to the vacuum system is<br />

∆k 2 = 0.031. Clearly the general services <strong>of</strong> the MSF unit are not affected by an<br />

intrinsic inefficiency but they have to consider product variation in order to account<br />

for its contribution to the final cost <strong>of</strong> water.<br />

The brine heater is located on the other side <strong>of</strong> the MSF plant. The effect <strong>of</strong> the<br />

inefficiency in the reject section also affects this component because the recycled<br />

brine heated in the brine heater comes from the reject section. The recycle brine<br />

flowing in the recovery section is reduced by 195 T/h <strong>and</strong> the cooling brine heating<br />

is reduced by 227 kW. The temperature difference in the first stage <strong>of</strong> the recovery<br />

section is improved by 0.1 ºC. So, heater efficiency decreases <strong>and</strong> the variation <strong>of</strong><br />

the unit exergy consumption is positive (∆k = 0.0044, see table A1.45). The<br />

malfunction is MF = 185 kW <strong>and</strong> an irreversibility increase in the heater <strong>of</strong><br />

∆I = 468 kW (see table A1.47).<br />

396 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.41 F-P values in design, NTOS case.<br />

Effect <strong>of</strong> reject section fouling<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 397


TABLE A1.42 F-P values when the fouling in RJS=0, NTOS case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

398 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.43 KP matrix in design, NTOS case.<br />

Effect <strong>of</strong> reject section fouling<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 399


TABLE A1.44 KP matrix with the inefficiency in RJS, NTOS case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

400 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.45 Variation <strong>of</strong> the KP matrix when the inefficiency in RJS is detected.<br />

Effect <strong>of</strong> reject section fouling<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 401


TABLE A1.46 Irreversibility matrix corresponding to reject fouling in RJS, NTOS case.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

402 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


TABLE A1.47 Dysfunction/malfunction table when the fouling in RJS=0.<br />

Effect <strong>of</strong> reject section fouling<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 403


FIGURE A1.25 Impact on fuel <strong>analysis</strong>, section A1.6.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

404 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

FIGURE A1.26 Increase <strong>of</strong> irreversibility in section A1.6.


Effect <strong>of</strong> reject section fouling<br />

As seen for the brine heater, the cleaning ball system in the reject section induces an<br />

unexpected 1,800 kW positive malfunction in the recovery section. This result will<br />

be described analytically. The temperature <strong>of</strong> water leaving the distiller is reduced<br />

by 1.7 ºC (remember that the cleaning ball system in the reject decreases the<br />

distillate pr<strong>of</strong>ile in the reject section <strong>and</strong>, therefore, in the last section <strong>of</strong> the recovery<br />

distiller). In general, since the heat transfer coefficient is higher at high<br />

temperatures, the thermal irreversibility increases in the recovery section<br />

(∆I = 2,079 kW, see table A1.47). As the product <strong>of</strong> the section decreases, the<br />

variation <strong>of</strong> the unit exergy consumption is positive (∆k = 0.223, see table A1.45).<br />

The impact on fuel associated with this induced malfunction is 4,248 kW.<br />

The malfunction associated with the fictitious device is –788 kW. Two fuels enter<br />

this component in the productive structure <strong>of</strong> the MSF unit, one is the exergy <strong>of</strong> the<br />

blowdown leaving the recovery section. This exergy is reduced because the<br />

temperature <strong>of</strong> the flashing brine decreases 1.8 ºC when leaving the reject section.<br />

So, the unit exergy consumption <strong>of</strong> the component is lower than in design<br />

(∆k = --0.017, see table A1.47). As demonstrated, a lower temperature <strong>of</strong> the<br />

blowdown rejected to the sea at least implies a lower cost in the water production.<br />

Finally, the mixer has an induced malfunction <strong>of</strong> 1,208 kW, with a very clear<br />

physical explanation. The temperatures <strong>of</strong> the make-up <strong>and</strong> flashing brine to<br />

blowdown are similar in the reference case but these temperatures are separated with<br />

the cleaning ball system in the reject section. The irreversibility generated in the<br />

mixing process is higher although those two flows are reduced to maintain the final<br />

production in the MSF plant (∆I = 1,191 kW, see table A1.47). The variation <strong>of</strong> the<br />

unit exergy consumption in the idealized component was ∆k = 0.0204 (see table<br />

A1.45). The additional fuel necessary for this component provoked by the cleaning<br />

ball system in RJS was 1,868 kW.<br />

In the dysfunction <strong>analysis</strong>, only the fictitious device had an important dysfunction<br />

generated by the inefficient components (total dysfunction was 3,283 kW). This<br />

component reduces its product by only 64 kW, however the final reduction in the<br />

distillate exergy flow is 482 kW.<br />

Although the plant diagnosis suggests that the MSF unit is working at a poorer<br />

efficiency (the impact on fuel associated with the unit exergy consumption variation<br />

was 6,394 kW), this <strong>analysis</strong> considered a constant total production. The<br />

temperature <strong>of</strong> the distillate leaving the MSF unit is 1.8 ºC lower than expected in<br />

design. This means that total production is not constant <strong>and</strong> the last term in equation<br />

(6.41) cannot be neglected. The impact on fuel associated with this variation is<br />

calculated by multiplying the total product variation by the exergy unit cost <strong>of</strong> the<br />

product. In this case 6,768 kW <strong>of</strong> fuel were saved (in the case <strong>of</strong> the power plant, the<br />

term <strong>of</strong> the product variation can usually be neglected because it is normally less<br />

than 20 kW). The total amount <strong>of</strong> fuel saved with this inefficiency is 374 kW, by<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 405


<strong>Thermoeconomic</strong> diagnosis<br />

combining the two effects. Therefore, the cleaning ball system also benefits the MSF<br />

unit, as well as the heater <strong>and</strong> recovery section.<br />

Figure A1.26 shows the effect <strong>of</strong> fouling in the reject section when we gradually<br />

decrease to zero the design value (0.000018 m 2 K/W). If the thermoeconomic model<br />

is linear with respect the variational <strong>analysis</strong> <strong>of</strong> the fouling, the malfunction matrix<br />

could be used to predict the impact on fuel associated with the desired variation <strong>of</strong><br />

the fouling <strong>of</strong> this component (if known).<br />

If the model responds linearly, the total cost <strong>of</strong> water (includes capital <strong>and</strong><br />

maintenance costs) must also increase linearly depending on the degree <strong>of</strong><br />

inefficiency (see figure A1.27). Each cubic meter <strong>of</strong> water increases 0.00012 $ when<br />

the fouling factor in the reject distiller increases 0.00001 m 2 K/W. Yearly freshwater<br />

production would involve an additional cost <strong>of</strong> 2,000 $ with this small variation in<br />

reject fouling.<br />

FIGURE A1.27 Effect on fuel consumption when the fouling in reject is varied. Nominal-temperature operation in<br />

summer (NTOS, i.e., 1,900 T/h <strong>and</strong> 32 ºC seawater temperature).<br />

-100<br />

-200<br />

-300<br />

-400<br />

-500<br />

fouling*10-5 in RJ<br />

0<br />

kW<br />

0 3 6 9 12 15 18<br />

Inc. fuel consumption<br />

The linearity <strong>of</strong> the model with respect to fouling variation is shown in figure A1.28.<br />

The malfunction matrix (table A1.48) can be used to predict the impact on fuel<br />

consumed with the inefficiency. But the induced malfunctions provoked by<br />

temperature variation in the rest <strong>of</strong> components implies that the <strong>analysis</strong> for several<br />

inefficiencies has different results than the individual <strong>analysis</strong> <strong>of</strong> those inefficiencies.<br />

So, the malfunction matrix can only be used to predict specific inefficiencies.<br />

Important errors may arise if it is used for several inefficiencies.<br />

The most important results derived from the <strong>analysis</strong> <strong>of</strong> the fouling in recovery<br />

section are :<br />

• Fouling increases the flash range <strong>of</strong> the plant <strong>and</strong>, therefore, the distillate<br />

production if the same control parameters <strong>of</strong> the plant are maintained. Input<br />

conditions must be relaxed to maintain the final production <strong>of</strong> freshwater.<br />

406 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Summary<br />

FIGURE A1.28 Variation <strong>of</strong> the water cost when fouling in the reject section is decreased from the design value<br />

to zero.<br />

1,474<br />

1,473<br />

1,472<br />

1,471<br />

$/m3<br />

• The inefficiency negatively affects the rest <strong>of</strong> the MSF components (see the<br />

malfunctions induced in other components in table A1.47). Furthermore, the<br />

benefit is due to the lower temperature <strong>of</strong> the freshwater produced, although the<br />

plant is not working more efficiently (the impact on fuel associated with the unit<br />

exergy consumption variation is positive). The cost <strong>of</strong> water is not reduced very<br />

much with the cleaning ball system.<br />

• The cleaning ball system is not recommended for the reject section. It is very<br />

difficult to install there (it is an open circuit in which some <strong>of</strong> the cooling brine is<br />

rejected to the sea), <strong>and</strong> the low temperatures do not provoke serious scaling<br />

problems in this section. Feed chlorination is a simpler solution to avoid possible<br />

biological fouling (which depends on seawater intake conditions).<br />

A1.7 Summary<br />

Water cost<br />

fouling*10-5 in RJ<br />

0 3 6 9 12 15 18<br />

<strong>Thermoeconomic</strong> diagnosis <strong>of</strong> the dual-purpose plant for the inefficiencies in<br />

section 7.3 was completed in this annex for the most representative load in the<br />

power <strong>and</strong> desalination plant. The symbolic formulation <strong>of</strong> the Structural Theory <strong>of</strong><br />

<strong>Thermoeconomic</strong>s provides a lot <strong>of</strong> information <strong>and</strong> explains the physical<br />

consequences expected with the inefficiency.<br />

Inefficiencies studied in steam power plant are local to the components suffering the<br />

inefficiency, but in the desalination plant the main units <strong>of</strong> the system are connected<br />

by the cooling brine, flashing brine <strong>and</strong> distillate, where any inefficiency is easily<br />

distributed over the rest <strong>of</strong> the plant components.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 407


TABLE A1.48 Malfunction matrix when the fouling in RJS is varied 0.00001 m 2 K/W.<br />

<strong>Thermoeconomic</strong> diagnosis<br />

408 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


ANNEX 2<br />

Thermodynamic properties<br />

<strong>of</strong> seawater<br />

Below are the models <strong>and</strong> correlations <strong>of</strong> the thermodynamic properties needed to<br />

simulate the MSF desalination plant, except for the properties previously described<br />

by the auxiliary equations (Chapter 3).<br />

A2.1 Specific enthalpy h <strong>of</strong> superheated or saturated vapor<br />

We used equations from Badr, Probert <strong>and</strong> O’Callaghan (1990), from formulations<br />

by Keenan <strong>and</strong> Keyes (1955, 1969) <strong>and</strong> conveniently expressed for computer<br />

calculation (Schnakel, 1958). The temperature <strong>and</strong> pressure range was valid below<br />

the critical point.<br />

Units: International System<br />

where<br />

p<br />

h F 101.31558 F0 ---------------------<br />

101325.0<br />

B0 ----- ⎛ p<br />

------------------------- ⎞<br />

2 ⎝101325.0T⎠ 2<br />

⎧<br />

= +<br />

⎨ +<br />

⎩<br />

⎛ p ⎞ ⎫<br />

– B6 + B0⎜B2– B3 + B0 B7 ------------------------- ⎟ ⎬<br />

⎝ 101325.0T ⎠ ⎭<br />

B0 p<br />

B B01 101325T 2<br />

----------------------- B2 – B ⎛<br />

B0 p<br />

3 --------------------- ⎞<br />

⎝101325T⎠ 2<br />

⎧ ⎫<br />

⎪ ⎪<br />

–<br />

⎨ +<br />

+ ( B4– B5) ⎬<br />

⎪ ⎪<br />

⎩ ⎭<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

2


410<br />

Thermodynamic properties <strong>of</strong> seawater<br />

B0<br />

= 1.89 – B1<br />

B2<br />

= 82.546<br />

B4<br />

= 0.21828 T<br />

B6<br />

= B0<br />

B3<br />

– 2 F0<br />

(B2<br />

– B3)<br />

B7<br />

= 2 F0<br />

(B4<br />

– B5)<br />

– B0<br />

B5<br />

2<br />

F = 1804036.3 + 1472.265 T + 0.37789824 T + 47845.137 ln T.<br />

A2.2 Specific entropy <strong>of</strong> superheated or saturated vapor<br />

Term ß was added to those in section A2.1. The specific entropy s <strong>of</strong> superheated<br />

vapor was:<br />

where<br />

B 1<br />

B 3<br />

B 5<br />

=<br />

=<br />

=<br />

s = 1472.626 ln T – 461.4874 ln p + 0.7557174 T + 3830.4065<br />

–<br />

2641.62 80870 ⁄ T2<br />

------------------ 10<br />

T<br />

162470<br />

-----------------<br />

T<br />

126970<br />

-----------------<br />

T<br />

372420<br />

F0 = 1.89 – B ⎛<br />

1 ----------------- + 2⎞<br />

⎝ ⎠<br />

β<br />

=<br />

T 2<br />

47845.076<br />

------------------------ – 101.31344 β<br />

T<br />

1<br />

p<br />

-- ( B0 – F0) -----------------<br />

T<br />

101325<br />

B ⎧<br />

0<br />

⎨<br />

+ -----<br />

⎩<br />

2<br />

⎫<br />

B0 ( B4 – B5) – 2B7 ⎬<br />

⎭<br />

⎛ p<br />

--------------------- ⎞<br />

⎝101325T⎠ 2 1<br />

B6 -- ⎛<br />

B0 p<br />

--------------------- ⎞<br />

2 ⎝101325T⎠ 2<br />

⎛ +<br />

⎞<br />

⎝ ⎠<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Specific volume <strong>of</strong> superheated or saturated vapor<br />

A2.3 Specific volume <strong>of</strong> superheated or saturated vapor<br />

Using B from section A2.1, the specific volume v <strong>of</strong> pure water was:<br />

A2.4 Latent heat vaporization <strong>of</strong> water as a function <strong>of</strong><br />

boiling temperature<br />

Below the atmospheric boiling point (373.15 K), latent heat <strong>of</strong> vaporization<br />

(SI units):<br />

where h was solved in section A2.1 <strong>and</strong> ps<br />

in section 3.3.6.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

λs<br />

was<br />

The Fish & Lielmezs correlation (Reid, Prausnitz <strong>and</strong> Sherwood, 1977) was used in<br />

the range 373.15 < T < 450 K:<br />

where<br />

⎛ T ⎞<br />

−3<br />

461539. 453<br />

v = 1. 00035⋅ 10 ⎜<br />

+ B⎟<br />

⎝ p ⎠<br />

λ s s<br />

= ( )<br />

( )<br />

h T, p ( T) −4. 186 T−273.<br />

15<br />

ℵ + ℵ0.35298<br />

λs 6051.1583<br />

1 ℵ 0.13856<br />

⎛ ⎞<br />

= ⎜------------------------------ ⎟ T<br />

⎝ + ⎠<br />

647.3 T<br />

ℵ 1.3615467 –<br />

= ⎛---------------------- ⎞<br />

⎝ T ⎠<br />

The Carruth & Kobayashi correlation (Reid et al., 1977) was used for<br />

450 < T < 647.3 K:<br />

λ s<br />

2115173.3 ⎛ T<br />

1 – ------------ ⎞<br />

⎝ 647.3⎠<br />

0.354<br />

1125343.9 ⎛ T<br />

1 – ------------ ⎞<br />

⎝ 647.3⎠<br />

0.456<br />

=<br />

+<br />

411


412<br />

Thermodynamic properties <strong>of</strong> seawater<br />

A2.5 Seawater exergy<br />

A2.5.1 Theory<br />

Mass flow <strong>and</strong> five parameter measurements characterize the different stages <strong>of</strong><br />

seawater: pressure, temperature, altitude, velocity <strong>and</strong> composition (Zaleta, Ranz<br />

<strong>and</strong> Valero, 1998). The exergy method associates each parameter with its exergetic<br />

component: mechanical, thermal, potential, kinetic <strong>and</strong> chemical, respectively.<br />

These components help to quantify some quality <strong>and</strong> quantity aspects <strong>of</strong> seawater.<br />

The information provided by the exergy method also clarifies concepts related to the<br />

seawater availability.<br />

Ambient reference<br />

The first step in developing the analytic exergy methodology is to establish the<br />

ambient reference (AR) for seawater comparison. The AR must be relatively<br />

abundant with respect to the rest <strong>of</strong> the systems or subsystems. The thermodynamic<br />

equilibrium conditions <strong>of</strong> AR must resemble a closed system; therefore, the system<br />

brought to AR conditions will undergo a series <strong>of</strong> physical-chemical changes.<br />

Authors sometimes call this the ‘dead state’, because it is a zero exergy state<br />

(although its energy is different than zero).<br />

AR may be chosen in different ways to establish thermodynamic equilibrium.<br />

Ahrendts (1980) proposes an approximation <strong>of</strong> the “dead” ambient <strong>of</strong> Earth if it<br />

were thermodynamically isolated from the rest <strong>of</strong> the universe. When we impose<br />

restrictions on the method (excluding HNO 3 formation <strong>and</strong> its products), the<br />

resulting AR composition is very similar to the real physical ambient. Liquid AR is<br />

mainly seawater with more than 99% <strong>of</strong> the system's total mass. On the other h<strong>and</strong>,<br />

Szargut (1980) proposes an AR that is more similar to the real physical ambient in<br />

nature <strong>and</strong> independent <strong>of</strong> the process or system under consideration. This is more<br />

convenient to exegetically analyze systems classified as natural resources.<br />

We used the AR proposed by Szargut to analyze seawater. The AR in the liquid<br />

phase corresponded to seawater composition at main ambient temperature <strong>and</strong> sea<br />

level atmospheric pressure. The seawater composition for the AR proposed by<br />

Szargut (1989) is shown in the next table.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Seawater exergy<br />

TABLE A2.1 Liquid phase composition <strong>of</strong> Reference Ambient (Szargut, 1989; Morris, <strong>and</strong> Szargut, 1986).<br />

Chemical element Molality (mol/kg)<br />

Ag (s) 2.7 × 10 –9<br />

As (s) 2.1 × 10 –8<br />

Au (s) 5.8 × 10 –11<br />

B (s) 3.4 × 10 –4<br />

Ba (s) 1.4 × 10 –7<br />

Bi (s) 1.0 × 10 –10<br />

Br 2 (l) 8.7 × 10 –4<br />

Ca (s) 9.6 × 10 –3<br />

Cd (s) 6.9 × 10 –11<br />

Cl2 (g) 0.5657<br />

Co (s) 6.8 × 10 –9<br />

Cs (s) 2.3 × 10 –9<br />

Cu (s) 7.3 × 10 –10<br />

F 2 (g) 3.87 × 10 –5<br />

Hg (l) 3.4 × 10 –10<br />

I 2 (s) 5.2 × 10 –7<br />

K (s) 1.04 × 10 –2<br />

Li (s) 2.5 × 10 –5<br />

Mg (s) 4.96 × 10 –2<br />

Mn (s) 7.5 × 10 –9<br />

Mo (s) 1.1 × 10 –7<br />

Na (s) 0.474<br />

Ni (s) 1.2 × 10 –7<br />

P (s) 4.9 × 10 –7<br />

Pb (s) 4.2 × 10 –11<br />

Rb (s) 1.42 × 10 –6<br />

S (s) 1.17 × 10 –2<br />

Se (s) 1.2 × 10 –9<br />

Sr (s) 8.7 × 10 –5<br />

W (s) 5.6 × 10 –10<br />

Zn (s) 1.7 × 10 –8<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 413


Thermodynamic properties <strong>of</strong> seawater<br />

Seawater availability: exergy function<br />

The availability <strong>of</strong> a renewable resource can be understood as ‘how accessible is it’.<br />

In order to be used, a resource must be changed chemically <strong>and</strong> physically to the<br />

required conditions (e.g., for human consumption, water must be extracted from a<br />

river or sea, be purified <strong>and</strong> sent to end users).<br />

The analogy between the availability <strong>of</strong> a natural resource <strong>and</strong> exergy helps relate<br />

each resource parameter with its exergy components. As the exergy method is<br />

conditioned by a Stable Reference Environment (SRE) —dead state conditions—<br />

the SRE proposed by Szargut (1980) is the most convenient (the most similar to the<br />

real physical environment <strong>of</strong> Earth).<br />

In the case <strong>of</strong> seawater, the exergy method is useful to quantify the ‘availability’ <strong>of</strong> a<br />

sea, with respect to the defined SRE. By applying the exergy model (Gaggioli, 1980)<br />

in terms <strong>of</strong> temperature, pressure, height, velocity <strong>and</strong> composition, <strong>and</strong> assuming<br />

seawater is an incompressible liquid <strong>and</strong> dilute substance, the specific exergy can be<br />

used in terms <strong>of</strong> its components for each seawater property (thermal, mechanical,<br />

chemical, kinetic <strong>and</strong> potential components, respectively):<br />

bar , CPH2O Ta Tr Tr Ln Ta =<br />

– -----<br />

Tr +<br />

∑<br />

i<br />

xir ,<br />

– vH2OPaP + ( – r)<br />

( µ ia , – µ ir , ) 1 2 2<br />

-- ( ca – cr ) + g( za– zr) 2<br />

414 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(A2.1)<br />

According to equation (A2.1), the thermal exergy component depends on the heat<br />

capacity <strong>of</strong> the aqueous solution <strong>and</strong> its absolute temperature T a. The mechanical<br />

exergy component is calculated from the specific volume <strong>of</strong> the solution (seawater)<br />

<strong>and</strong> the pressure difference between the sea <strong>and</strong> the SRE. The specific heating value<br />

CP H2 O <strong>and</strong> the specific volume v H2O <strong>of</strong> the solution can be calculated without<br />

serious error if it is considered pure water (Perry <strong>and</strong> Chilton, 1984). We used the<br />

correlations described in Chapter 4. The potential exergy component requires the<br />

altitude z above sea level (almost negligible in a MSF plant). It is used to calculate<br />

the maximum mechanical work obtained from a waterfall, such as a hydroelectric<br />

station. The kinetic exergy component is <strong>of</strong> relatively little exergetic importance in<br />

comparison with other exergetic components (taking into account the low velocity c a<br />

<strong>of</strong> brine inside the tubes or in the flash chambers). Its mean velocity must be<br />

calculated, which depends on flow <strong>and</strong> operation conditions. The chemical exergy<br />

component is the most complex to calculate. It may be broken down into the<br />

following components: (i) the chemical exergy <strong>of</strong> the water, (ii) the chemical exergy<br />

<strong>of</strong> the dissolved inorganic substances, (iii) the chemical exergy <strong>of</strong> the organic<br />

substances.


Seawater exergy<br />

i) The chemical exergy <strong>of</strong> pure water in the sea. This component provides information<br />

about the thermodynamic degradation process; pure water availability under<br />

different conditions such as pollution (the presence <strong>of</strong> substances other than pure<br />

water like salts, organic material, etc.). The magnitude <strong>of</strong> the exergetic component<br />

µ can be calculated from its activity as a pure substance in a solution equation<br />

(equation A2.2):<br />

, xH2O µ H2O µ ( – H2Or , ) xH2O RTr Ln<br />

⎛ ⎞<br />

= =<br />

⎜-------------- ⎟<br />

⎝aH2O, r⎠<br />

b qH2O<br />

(A2.2)<br />

where x H2O is the molar fraction <strong>of</strong> pure water in seawater, <strong>and</strong> a H2O, a H2O,r can<br />

be estimated from measuring coligative properties, such as osmotic pressure, π.<br />

In the case <strong>of</strong> seawater, the osmotic pressure <strong>of</strong> a diluted solution with respect to<br />

its pure solvent is typically calculated using equation A2.3,<br />

π H2O<br />

= – -------- Ln ( a <strong>and</strong> (A2.3)<br />

v H2O)<br />

π H2O, r = – -------- Ln ( a<br />

v H2O, r)<br />

where π is obtained by measuring the Electrical Conductivity (EC) <strong>of</strong> seawater<br />

<strong>and</strong> considering that the osmotic pressure is a function <strong>of</strong> the salt concentration<br />

(binary) in solution (without any serious errors, as in the case <strong>of</strong> a very diluted<br />

substance, such as seawater).<br />

π H2 O<br />

= 0.36 EC (A2.4)<br />

where π is the osmotic pressure (atmospheres) <strong>and</strong> EC the electrical conductivity<br />

in dS/m (1 dS/m = 640 ppm, Medina (2000)) <strong>of</strong> ionized electrolytic components<br />

in a solution.<br />

ii) The chemical exergy <strong>of</strong> the dissolved inorganic substance is determined by the<br />

well-known procedure for an electrolytic solution (equation A2.5):<br />

b qi<br />

(A2.5)<br />

where the activity for each chemical substance i in the sea <strong>and</strong> in the SRE can be<br />

expressed in terms <strong>of</strong> the activity coefficient, γ, <strong>and</strong> its molality, m:<br />

a i = γ i m i<br />

RT r<br />

, = xi ( µ i – µ ir , ) =<br />

xi RTr Ln ⎛------- ⎞<br />

⎝ ⎠<br />

RT r<br />

a H2O<br />

(A2.6)<br />

The activity coefficient, γ, <strong>of</strong> each <strong>of</strong> the electrolytic species is determined using<br />

the equation obtained by Debye-Hückel (equation A2.7).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 415<br />

a i<br />

air ,


Thermodynamic properties <strong>of</strong> seawater<br />

Log( γ i)<br />

416 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

(A2.7)<br />

where A, B are constants depending on the solvent <strong>and</strong> temperature, z i is the ionic<br />

charge, d i is the ionic diameter <strong>of</strong> specie i <strong>and</strong> I is the ionic dissolution force,<br />

∑<br />

2<br />

I = mizi. For diluted solutions (seawater is a good example), this equation can<br />

i<br />

be expressed as:<br />

=<br />

–<br />

1 2<br />

Azi I ⁄<br />

1 2<br />

1+ BdiI ⁄<br />

----------------------------<br />

1 2 ⁄<br />

Log ( γ i)<br />

= – [ Azi I ]<br />

(A2.8)<br />

The activity coefficient <strong>of</strong> non-electrolytic inorganic substances is always γ =1.<br />

iii) The chemical exergy <strong>of</strong> organic substances. It is difficult to determine the presence<br />

<strong>of</strong> organic substances in seawater due to the diversity <strong>of</strong> species involved<br />

(including biological organisms). However, organic substances are not present in<br />

the Szargut (1980) definition <strong>of</strong> SRE, but are purified naturally in rivers. This<br />

means that the oxygen (from photosynthesis or atmospheric) dissolved in water<br />

oxidizes the organic substances. This process may be slow or fast depending on<br />

the substance. One way to quantify the exergetic content <strong>of</strong> an organic substance<br />

is by proposing a single organic molecule to represent the “organic substance<br />

mean”.<br />

For practical sea <strong>analysis</strong>, our representative substance was a fat molecule, as shown<br />

in equation A2.9. This enabled us to calculate the order <strong>of</strong> magnitude <strong>of</strong> the exergy<br />

organic component to be determined qualitatively.<br />

115<br />

C39 H80 O3 + -------- O2 ↔ 39 CO2 + 40 H2O 2<br />

(A2.9)<br />

The laboratory measurement <strong>of</strong> Chemical Oxygen Dem<strong>and</strong> (COD, mg. <strong>of</strong> O 2/lt <strong>of</strong><br />

seawater consumed in the reaction is estimated) was used to obtain the amount <strong>of</strong><br />

moles <strong>of</strong> mean organic substance per liter <strong>of</strong> water. The exergy <strong>of</strong> the organic<br />

substance was obtained from the definition <strong>of</strong> exergy reaction in the st<strong>and</strong>ard state,<br />

according to the expression in equation A2.10.<br />

b o<br />

o<br />

∆hf =<br />

∑<br />

o o o<br />

o<br />

∆hf – T s – xj µ j<br />

o<br />

µ j<br />

where , s o <strong>and</strong> are well known for industrial substances.<br />

(A2.10)


Seawater exergy<br />

A2.5.2 Practice: Brine exergy as a function <strong>of</strong> temperature, pressure<br />

<strong>and</strong> salt concentration<br />

Brine exergy only includes thermal, chemical <strong>and</strong> mechanical terms (kinetic <strong>and</strong><br />

potential terms are neglected, see equation A2.1). Although it is impossible to know<br />

the chemical <strong>analysis</strong> <strong>of</strong> seawater entering the MSF unit, the chemical term only<br />

considers seawater concentration due to sodium chloride.<br />

This means that the chemical energy <strong>of</strong> the organic compounds is not considered<br />

<strong>and</strong> the contribution <strong>of</strong> inorganic substances is only calculated for Na + <strong>and</strong> Cl – ions.<br />

Chemical exergy <strong>of</strong> pure water depends on the osmotic pressure difference with<br />

respect to reference seawater. The AR used was 0 ºC <strong>and</strong> 45,000 TDS (average<br />

seawater concentration in the Arabian Gulf). The results were similar to other<br />

studies (Zaleta et al., 1998). For more detailed information about how to calculate<br />

these terms, see Barner <strong>and</strong> Scheuerman (1978), Newman (1980) <strong>and</strong> Marín <strong>and</strong><br />

Turégano (1985).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 417


ANNEX 3<br />

Technical data<br />

This annex includes the most important design <strong>and</strong> constructive values provided by<br />

the contractors. Most <strong>of</strong> those values are introduced in the simulator, but they cannot<br />

be changed unless requested by the author.<br />

A3.1 MSF plant<br />

MSF: Guarantee figures (112 ºC TBT, 25 ºC SWT)<br />

Seawater temperature (Tsea)<br />

25 (ºC)<br />

Distillate production per hour (D) 2,400 (T/h)<br />

Distillate temperature at pump suction 38 (ºC)<br />

3<br />

Distillate density at production temperature 994 (kg/m )<br />

Discharge pressure at distillate pump 3.5 (bar)<br />

Distillate purity expressed as TDS 10 (ppm)<br />

pH value <strong>of</strong> distillate before caustic soda injection 5.5-6.0<br />

Fe content in distillate 0.05 (ppm)<br />

Cu content in distillate 0.05 (ppm)<br />

Vapor velocity at the smallest path in last stage 14 (m/s)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


420<br />

Technical data<br />

Performance ratio (PR) not less than 8<br />

Quantity <strong>of</strong> heating steam at reducing valve before brine heater (mST)<br />

313.400 (kg/h)<br />

Steam pressure at heater inlet 1.8 (bar)<br />

Steam temperature at heater inlet 120 (ºC)<br />

Heater condensate temperature at pump suction 117 (ºC)<br />

Net specific heat consumption (NC) 290.75 (kJ/t distillate)<br />

Total specific heat consumption 295 (kJ/kg distillate)<br />

–3<br />

Specific electric power consumption 4.0 (kWh/kg dist. x 10 )<br />

O2<br />

content in heater condensate (at 20 ºC) 0.03 (ppm)<br />

Fe content in heater condensate 0.04 (ppm)<br />

Cu content in heater condensate 0.04 (ppm)<br />

Conductivity <strong>of</strong> heater condensate 5 (µs/cm)<br />

Temperature <strong>of</strong> ejector condensate 40 (ºC)<br />

PH <strong>of</strong> ejector condensate 5.5-6.0<br />

T.D.S in brine blow down 71,000 (ppm, máx.)<br />

T.D.S in recirculated brine in the heater tubes 62,000 (ppm)<br />

Temperature <strong>of</strong> the sea water outlet from heat rejection section 36 (ºC)<br />

Sea water velocity inside tubes <strong>of</strong> heat rejection section 2.0 (m/s)<br />

Brine velocity inside tubes <strong>of</strong> heat recovery section 2.1 (m/s)<br />

Brine velocity inside tubes <strong>of</strong> brine heater 2.1 (m/s)<br />

Pressure inside the heater space 1.8 (bar)<br />

Brine pressure after the heater 1.9 (bar)<br />

Brine temperature in first stage (TBT) 108 (ºC)<br />

Brine temperature in last stage 35.5 (ºC)<br />

Vapor temperature in first stage 106.5 (ºC)<br />

Vapor temperature in last stage 34.5 (ºC)<br />

Absolute pressure in first stage 1.305 (bar)<br />

Absolute pressure in last stage 0.055 (bar)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


MSF plant<br />

Temperature <strong>of</strong> make-up feed entering deaerator 36 (ºC)<br />

Temperature <strong>of</strong> make-up feed leaving deaerator 36 (ºC)<br />

Absolute pressure in deaerator space 0.05 (bar)<br />

O2<br />

content in feed make-up leaving deaerator (without sulphite inj.) 0.03 (ppm)<br />

O2<br />

content in feed make-up leaving deaerator (with sulphite inject.) 0.04 (ppm)<br />

–6<br />

Specific chemical consumption (antiscale with sponge ball cleaning) 12 (kg/kg dist. x 10 )<br />

–6<br />

Specific chemical consumption (antiscale without sponge ball clean.) 27.2 (kg/kg dist. x 10 )<br />

7<br />

Heat losses due to radiation, venting or other losses 5 x 10 (kJ/h)<br />

Evaporators<br />

GENERAL<br />

2<br />

Recovery section: heat exchange surface 110,200 (m )<br />

2<br />

Reject section: heat exchange surface 15,150 (m )<br />

2<br />

Brine heater: heat exchange surface 10,272 (m )<br />

2<br />

Recovery section: Fouling factor (design) 0.00015 (m K/W)<br />

2<br />

Reject section: Fouling factor (design) 0.00018 (m K/W)<br />

2<br />

Brine heater: Fouling factor (design) 0.00025 (m K/W)<br />

2<br />

Recovery section: Heat transfer coefficient (design) 2,673 (W/m K)<br />

2<br />

Reject section: Heat transfer coefficient (design) 2,211 (W/m K)<br />

2<br />

Brine heater: Heat transfer coefficient (design) 2,147 (W/m K)<br />

2<br />

Demisters: Total area 640 (m )<br />

Total width 19 (m)<br />

Total length 87 (m)<br />

Total height 17 (m)<br />

Total weight-empty 3,000,000 (kg)<br />

Tube Pitch (pattern: triangular) 1.25<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

421


422<br />

Technical data<br />

BRINE HEATER<br />

Number <strong>of</strong> tubes 3060 (2 passes)<br />

Tube internal diameter 33 (mm)<br />

Tube thickness 1.2 (mm)<br />

Tube effective length 15.1 (m)<br />

Tube material CuNi 66/30 2 Fe 2 Mn<br />

Tube conductivity 28.0 (W/m K)<br />

RECOVERY SECTION: Stages 1-2<br />

Number <strong>of</strong> tubes 3060<br />

Tube internal diameter 33 (mm)<br />

Tube thickness 1.0 (mm)<br />

Tube effective length 19.2 (m)<br />

Tube material CuNi 70/30 ASTM B107<br />

Tube conductivity 31.1 (W/m K)<br />

RECOVERY SECTION: Stages 3-11<br />

Number <strong>of</strong> tubes 3060<br />

Tube internal diameter 33 (mm)<br />

Tube thickness 1.2 (mm)<br />

Tube effective length 19.2 (m)<br />

Tube material CuNi 90/10 ASTM B111<br />

Tube conductivity 51.9 (W/m K)<br />

RECOVERY SECTION: Stages 12-17<br />

Number <strong>of</strong> tubes 3185<br />

Tube internal diameter 33 (mm)<br />

Tube thickness 0.5 (mm)<br />

Tube effective length 19.2 (m)<br />

Tube material Titanium B338Gr2<br />

Tube conductivity 22.0 (W/m K)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


MSF plant<br />

REJECT SECTION: Stages 18-20<br />

Number <strong>of</strong> tubes 2390<br />

Tube internal diameter 33.6 (mm.)<br />

Tube thickness 0.7<br />

Tube effective length 19.2<br />

Tube material Titanium B338Gr2<br />

Tube conductivity 22.0 (W/m K)<br />

EFFECTIVE STAGE LENGTHS AND WIDTHS FOR BRINE FLOW<br />

Stage no. Length (m) Width (m)<br />

1 3.800 19.000<br />

2 3.800 19.000<br />

3 3.800 19.000<br />

4 3.800 19.000<br />

5 3.800 19.000<br />

6 4.000 19.000<br />

7 4.000 19.000<br />

8 4.000 19.000<br />

9 4.200 19.000<br />

10 4.200 19.000<br />

11 4.400 17.500<br />

12 4.400 17.500<br />

13 4.500 17.500<br />

14 4.500 17.500<br />

15 4.800 17.500<br />

16 4.800 17.500<br />

17 4.000 17.500<br />

18 4.800 17.500<br />

19 4.300 17.500<br />

20 5.100 17.500<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

423


424<br />

Technical data<br />

DEMISTERS<br />

Stage no. 2 Area (m )<br />

Height (m)<br />

1 26.89 2.8<br />

2 22.00 2.8<br />

3 22.00 2.8<br />

4 22.00 2.8<br />

5 22.00 2.8<br />

6 25.75 2.8<br />

7 25.75 2.8<br />

8 25.75 2.8<br />

9 29.50 2.8<br />

10 29.50 2.8<br />

11 33.30 2.8<br />

12 33.30 2.8<br />

13 35.20 2.8<br />

14 35.20 2.8<br />

15 40.80 2.8<br />

16 40.80 2.8<br />

17 40.80 2.8<br />

18 35.10 2.8<br />

19 38.80 2.8<br />

20 52.33 2.8<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


MSF plant<br />

BRINE ORIFICES (112 ºC TBT, 25 ºC SW)<br />

Stage no. Height (mm) Width (mm) 3 Area (m )<br />

1 77 16.134 19.000<br />

2 80 16.134 19.000<br />

3 83 16.134 19.000<br />

4 87 16.134 19.000<br />

5 91 16.134 19.000<br />

6 95 16.134 19.000<br />

7 99 16.134 19.000<br />

8 104 16.134 19.000<br />

9 108 16.134 19.000<br />

10 113 16.134 19.000<br />

11 131 14.420 17.500<br />

12 137 14.420 17.500<br />

13 144 14.420 17.500<br />

14 150 14.420 17.500<br />

15 156 14.420 17.500<br />

16 163 14.420 17.500<br />

17 169 14.420 17.500<br />

18 175 14.420 17.500<br />

19 182 14.420 17.500<br />

20 200 14.420 17.500<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

425


426<br />

Technical data<br />

A3.2 Power Plant<br />

Boiler<br />

GENERAL<br />

Length x width x height (furnace) 9.825 x 10.875 x 19.9 (m)<br />

Length x width x height (steel structure) 23.0 x 15.5 x 45.5 (m)<br />

Total weight <strong>of</strong> boiler unit 3,500 (T)<br />

Shipping volume <strong>of</strong> largest item 3<br />

120 (m )<br />

Total gross weight <strong>of</strong> the largest item to be shipped 80 (T)<br />

Weight <strong>of</strong> the largest item to be dismantled during maintenance 15 (T)<br />

ECONOMIZERS<br />

Effective heating surface (ECO 1/ ECO 2) 2<br />

10,890/4,390 (m )<br />

Number <strong>of</strong> stages in line (ECO 1/ ECO 2) 7/3<br />

Number <strong>of</strong> parallel streams (ECO 1/ ECO 2) 1/1<br />

Location (ECO 1/ ECO 2) rd rd nd<br />

3 /3 -2 pass<br />

Design pressure 129 (bar)<br />

Design temperature (ECO 1/ ECO 2) 260/355 (ºC)<br />

Effective height <strong>of</strong> one stage 1,555 (mm)<br />

Pitch across the gas flow (ECO 1/ ECO 2) 65/75 (mm)<br />

Pitch parallel to the gas flow 75/110 (mm)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Power Plant<br />

AIR WATER HEATER<br />

Number <strong>of</strong> heaters per boiler 2<br />

Design pressure (airside) 1,300 (mm WG)<br />

Design pressure (waterside) 129 (bar)<br />

Design temperature (airside) 250 (º C)<br />

Design temperature (waterside) 260 (º C)<br />

Design air throughput 3<br />

463,740 (Nm /h)<br />

Design water throughput 211 (t/h)<br />

Effective surface heating 2<br />

20,920 (m )<br />

Fouling factor considered (air/water side) 5/2 %<br />

STEAM WATER DRUM<br />

Type 3<br />

48 (m )<br />

Water content 3<br />

24 (m /h)<br />

Steam space rating 3 3<br />

470 (m /m ·h)<br />

Design pressure 129 (bar)<br />

Design temperature 330 (º C)<br />

Total length 14,000 (mm)<br />

Shell length 12,800 (mm)<br />

Shell thickness 82 (mm)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

427


428<br />

Technical data<br />

Combustion chamber<br />

WALL HEATING SURFACES<br />

Nominal height 19.9 (m)<br />

Nominal width 10.875 (m)<br />

Nominal depth 9.825 (m)<br />

Volume 3<br />

2.123 (m )<br />

Total effective heat absorbing surface <strong>of</strong> the combustion chamber 2<br />

1,454 (m )<br />

Total length 14,000 (mm)<br />

Shell length 12,800 (mm)<br />

Shell thickness 82 (mm)<br />

Heat input (natural gas at MCR, 40º C air temperature) 6<br />

422.22 × 10 (kcal/h)<br />

Evaporators<br />

Total effective heat absorbing surface 2<br />

2,740 (m )<br />

Design pressure 129 (bar)<br />

Design temperature 375 (ºC)<br />

Maximum local heat flux 2<br />

290,000 (kcal/m ·h)<br />

Evaporator headers<br />

Number 40<br />

Design pressure 129 (bar)<br />

Design temperature 330 (ºC)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Power Plant<br />

SUPERHEATERS<br />

Number <strong>of</strong> stages in line 3<br />

Number <strong>of</strong> parallel streams 2<br />

Number <strong>of</strong> spray attemperators 4<br />

Design pressure 129 (bar)<br />

Design temperature (máx.) (SH1/SH2/SH3) 580/590/590 (ºC)<br />

Effective heating surface (SH1/SH2/SH3) 2<br />

3,090/860/360 (m )<br />

Number <strong>of</strong> elements over the width (SH1/SH2/SH3) 144/72/72<br />

SPRAY ATTEMPERATORS<br />

Number 2<br />

Design steam flow (inlet/outlet) (AT1/AT2) 270-295/295-310 (t/h)<br />

Calculated spray water flow (AT1/AT2) 27/18 (t/h)<br />

Design spray water flow (AT1/AT2) 41/27 (t/h)<br />

Design pressure 129 (bar)<br />

Design temperature (AT1/AT2) 500/550 (ºC)<br />

DOWNCOMERS<br />

Number 2<br />

Outside diameter 508 (mm)<br />

Wall thickness 16 (mm)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

429


430<br />

Technical data<br />

Condensing Plant<br />

Condenser surface (between tube sheets <strong>and</strong> related to steam side) 2<br />

6,725 (m )<br />

Condenser vacuum at MCR 0.072 (bar abs)<br />

Specific condenser surface dem<strong>and</strong> at MCR 2 67.5 (m ·h/t)<br />

Condenser hotwell useful capacity 3 25 (m )<br />

Circulating water velocity within tube bundle 2.2 (m/s)<br />

Associated hydraulic loss <strong>of</strong> CW 0.37 (bar)<br />

Basic heat transfer coefficient at MCR 2<br />

2,732 (kcal/m ·h·K)<br />

Applied cleanliness factor 90 %<br />

Associated maximum temperature difference 6.7 (ºC)<br />

Thermal conductivity 14 (kcal/m·h·K)<br />

Number <strong>of</strong> tubes per total cond. for one turbine 7124<br />

Condensate Pumps<br />

Number <strong>of</strong> pumps 2 + 2<br />

Specific gravity <strong>of</strong> fluid (MCR) 3<br />

992.5 (kg/m )<br />

Suction pressure (MCR) 0.071 (bar)<br />

Suction temperature (MCR) 39.2 (ºC)<br />

Discharge pressure (MCR) 18 (bar abs.)<br />

Discharge temperature (MCR) 39.2 (ºC)<br />

Flow at discharge nozzle (MCR) 2 x 131 (T/h)<br />

Overall efficiency according to DIN 1944 <strong>of</strong> equiv. (MCR) 71.6 %<br />

Pump speed 1485 (l/min)<br />

Critical speeds <strong>of</strong> pump <strong>and</strong> motor unit > 1800 (rpm)<br />

Nameplate rating (MCR) 130 (kW)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Power Plant<br />

HP1 Heater<br />

Overall dimensions <strong>of</strong> feed heater 1,300 x 8,600 (mm)<br />

Main steam flow (feedwater side, MCR) 562.9 (t/h)<br />

Inlet pressure (feedwater side, MCR) 119.05 (bar)<br />

Inlet/outlet temperature (feedwater side, MCR) 194.6/230.1 (ºC)<br />

Heating steam flow 39.0 (t/h)<br />

Pressure incl. vacuum if appl. 27.2 (bar)<br />

Temperature (heating side, MCR) 369 (ºC)<br />

Applied cleanliness factor 80 %<br />

Overall heat transfer coefficient (condensing zone) 2<br />

3,280 (kcal/m ·h·K)<br />

LMTD (condensing zone) 11.6 (ºC)<br />

Heat transfer surface (desuperheating section) 2<br />

65.3 (m )<br />

Heat transfer surface (condensing section) 2<br />

531.6 (m )<br />

Heat transfer surface (condensate cooling section) 2<br />

64.4 (m )<br />

Velocity <strong>of</strong> main condensate or feed water inside tubes 1.54 (m/s)<br />

HP2 Heater<br />

Overall dimensions <strong>of</strong> feed heater 1,300 x 8,600 (mm)<br />

Main steam flow (feedwater side, MCR) 562.3 (t/h)<br />

Inlet pressure (feedwater side, MCR) 119.4 (bar)<br />

Inlet/outlet temperature (feedwater side, MCR) 164.8/194.6 (ºC)<br />

Heating steam flow 29.9 (t/h)<br />

Pressure incl. vacuum if appl. 14.12 (bar)<br />

Temperature (heating side, MCR) 282 (ºC)<br />

Applied cleanliness factor 80 %<br />

Overall heat transfer coefficient (condensing zone) 2<br />

3,200 (kcal/m ·h·K)<br />

LMTD (condensing zone) 10.54 (ºC)<br />

Heat transfer surface (desuperheating section) 2<br />

37.8 (m )<br />

Heat transfer surface (condensing section) 2<br />

525.2 (m )<br />

Heat transfer surface (condensate cooling section) 2<br />

101.3 (m )<br />

Velocity <strong>of</strong> main condensate or feed water inside tubes 1.48 (m/s)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

431


432<br />

Technical data<br />

LP1 Heater<br />

Overall dimensions <strong>of</strong> feed heater 1,124 x 8,800 (mm)<br />

Main steam flow (feedwater side MCR) 131.6 (t/h)<br />

Inlet pressure (feedwater side, MCR) 11.622 (bar)<br />

Inlet/outlet temperature (feedwater side, MCR) 78.2/128.2 (ºC)<br />

Heating steam flow 12.0 (t/h)<br />

Pressure incl. vacuum if appl. 2.7 (bar)<br />

Temperature (heating side, MCR) 129.7 (ºC)<br />

Applied cleanliness factor 80 %<br />

Overall heat transfer coefficient (condensing zone) 2<br />

3,200 (kcal/m ·h·K)<br />

LMTD (condensing zone) 23.856 (ºC)<br />

Heat transfer surface (condensing section) 2<br />

341.4 (m )<br />

Heat transfer surface (condensate cooling section) 2<br />

54.1 (m )<br />

Velocity <strong>of</strong> main condensate or feed water inside tubes 1.76 (m/s)<br />

LP2 Heater<br />

Overall dimensions <strong>of</strong> feed heater 1,124 x 9,900 (mm)<br />

Main steam flow (feedwater side MCR) 131.6 (t/h)<br />

Inlet pressure (feedwater side, MCR) 12.072 (bar)<br />

Inlet/outlet temperature (feedwater side, MCR) 41.0/78.2 (ºC)<br />

Heating steam flow 8.2 (t/h)<br />

Pressure incl. vacuum if appl. 0.47 (bar)<br />

Temperature (heating side, MCR) 79.7 (ºC)<br />

Applied cleanliness factor 80 %<br />

Overall heat transfer coefficient (condensing zone) 2<br />

2,840 (kcal/m ·h·K)<br />

LMTD (condensing zone) 22.89 (ºC)<br />

Heat transfer surface (condensing section) 2<br />

316.4 (m )<br />

Heat transfer surface (condensate cooling section) 2<br />

124.8 (m )<br />

Velocity <strong>of</strong> main condensate or feed water inside tubes 1.61 (m/s)<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Nomenclature<br />

Abbreviatures/ Symbols/Acronyms<br />

a Cost parameter, activity, or constant <strong>of</strong> Cobb-Douglass equation.<br />

A Exchange area <strong>of</strong> the evaporator/condenser or constant <strong>of</strong> Debye-Hückel<br />

equation.<br />

AR Reference Ambient.<br />

AT Atemperator.<br />

b Specific exergy.<br />

B Flashing brine flow in j-th flash chamber, exergy flow, constant <strong>of</strong> Debye-<br />

Hückel equation, or constant for calculating vapor enthalpy.<br />

BD Brine Blowdown.<br />

BDP Blowdown Pump.<br />

BH Brine Heater.<br />

BHP Brine Heater Pump.<br />

BOI Boiler.<br />

BPE Boiling Point Elevation <strong>of</strong> brine with respect the pure water.<br />

c Velocity.<br />

C Salt concentration, or total monetary cost.<br />

c* Exergoeconomic cost.<br />

ca Cost per unit <strong>of</strong> area.<br />

CBS Cleaning Ball System.<br />

cf Fuel cost.<br />

CND Condenser.<br />

COC Boiler Peak Load.<br />

COD Chemical Oxygen Dem<strong>and</strong>.<br />

CP Condensate Pump or Heat Capacity.<br />

cp Product cost.<br />

CW Cooling rejected Water.<br />

d Ionic diameter.<br />

D Distillate flow.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


434<br />

Nomenclature<br />

DAS Data Acquisition System.<br />

DB Exergy flow <strong>of</strong> distillate.<br />

DCA Drain Cooling Advantage.<br />

DF Dysfunction generated in a component.<br />

DI Dysfunction generated by a component.<br />

DLL Dynamic Link Library.<br />

DP Distillate Pump.<br />

DRT Deaerator.<br />

DV Main stop valve seat diameter.<br />

e Condenser efficiency.<br />

E Enhancement factor.<br />

EC Electrical Conductivity.<br />

ECO Economizer.<br />

ED Electrodyalisis.<br />

EDS European Desalination Society.<br />

EES Engineering Equation Solver.<br />

ESL Excitation System Losses.<br />

f Generic function.<br />

F Fuel, Make-up feed or constant for calculating vapor enthalpy.<br />

FCW Fuel Cost <strong>of</strong> Water.<br />

FD Fictitious Device.<br />

FP Feed Pump.<br />

g Acceleration due to gravity, or characteristic equation.<br />

Gc Gas consumption.<br />

GCC Gulf Council Countries.<br />

GEN Generator.<br />

GOR Gain Output Ratio.<br />

h Heat transfer coefficient or enthalpy.<br />

H Height.<br />

Hb Flashing brine (seawater) enthalpy.<br />

HHV High Heating Value.<br />

HT High-Temperature.<br />

HP High-Pressure.<br />

HPH High-Pressure Heater.<br />

HPT High-Pressure Turbine.<br />

HR Heat Rate <strong>of</strong> a power plant.<br />

HRSG Heat Recovery Steam Generator.<br />

HTOS High-Temperature Operation in Summer.<br />

HTOW High-Temperature Operation in Winter.<br />

Hv Saturated vapor enthalpy <strong>of</strong> water.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Nomenclature<br />

I Irreversibility or Ionic dissolution force.<br />

IAAE International Agency <strong>of</strong> Atomic Energy.<br />

ID Inside Diameter.<br />

IDA International Desalination Association.<br />

k Thermal conductivity or unit exergy consumption.<br />

K Constant for mass flow coefficient or gl<strong>and</strong> steam system.<br />

k* Exergy unit cost.<br />

L Length or Exergy Losses.<br />

LP Low-Pressure.<br />

LPH Low-Pressure Heater.<br />

LPT Low-Pressure Turbine.<br />

LS Live Steam Extraction.<br />

LTL Low Turbine Load.<br />

LTMD Logarithmic Temperature Mean Difference.<br />

LTOS Low-Temperature Operation in Summer.<br />

m Mass flow or molality.<br />

MCR Maximum Continuous Rating.<br />

Md Steam flow to MSF unit.<br />

MED Multi-Effect Distillation.<br />

MF Malfunction <strong>of</strong> a component.<br />

MF* Malfunction cost (impact on fuel).<br />

MFl Intrinsic malfunction.<br />

MFg Induced malfunction.<br />

MIX Mixer.<br />

MR Maximum Rating.<br />

MSL Minimum Stable Load.<br />

MSF Multistage Flash.<br />

MXT Mixer Temper water.<br />

n number <strong>of</strong> tubes in a vertical row.<br />

NC Net energy Consumption.<br />

NEA Non Equilibrium Allowance.<br />

NRC Number <strong>of</strong> Recovery Stages.<br />

NRJ Number <strong>of</strong> Reject Stages.<br />

NTL Normal Turbine Load.<br />

NTOS Nominal-Temperature Operation in Summer.<br />

NTW Non Turbine Working.<br />

OD Outside Diameter.<br />

ODOB One Desalination One Boiler.<br />

O&M Operating <strong>and</strong> Maintenance<br />

p Pressure.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 435


436<br />

Nomenclature<br />

P Product.<br />

Pc Condenser pressure.<br />

Pr Pr<strong>and</strong>tl number.<br />

PE Pressure Exchanger.<br />

PL Pressure losses, or Partial Load.<br />

PR Performance Ratio.<br />

PTC Performance Test Case or Parabolic Trough Collector.<br />

Q Heat flow.<br />

Qf Heat value <strong>of</strong> fuel.<br />

r Exergy ratio.<br />

R Thermal resistance or recycle brine.<br />

RCS Recovery Section.<br />

Re Reynolds number.<br />

RJS Reject Section.<br />

RO Reverse Osmosis.<br />

rp Pressure ratio in a turbine section.<br />

RP Recycle Pump.<br />

s Specific entropy.<br />

S Entropy flow or size.<br />

Sa Sonic area.<br />

SF Solar Factor.<br />

SH Superheater.<br />

SR Seawater to Reject section flow.<br />

SRE Stable Reference Environment<br />

SW Seawater feed flow.<br />

SWP Seawater Pump.<br />

SWRO Seawater Reverse Osmosis.<br />

t Thickness.<br />

T Temperature.<br />

T* Temperature reference, 273.15 K.<br />

TBT Top Brine Temperature.<br />

TDOB Two Desalination One Boiler.<br />

TDS Total Dissolved Solids.<br />

To Ambient Temperature.<br />

TP Temper water Pump (also TPP).<br />

TTD Terminal Temperature Difference.<br />

TVC Thermal Vapor Compression.<br />

UAE United Arab Emirates.<br />

U Overall heat transfer coefficient.<br />

US, USA United States <strong>of</strong> America.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Nomenclature<br />

VC Vapor Compression.<br />

VEX Extraction valve (pressure loss <strong>simulation</strong>).<br />

VF Feed valve.<br />

vw Tube velocity.<br />

VS Reducing pressure station valve.<br />

VST Stop valve.<br />

VTE Vertical Tube Evaporator.<br />

VWO Valve Wide Open.<br />

x<br />

Variable or molar fraction.<br />

X Steam quality.<br />

w Width.<br />

W Power.<br />

z Ionic charge.<br />

Z Pressure drop coefficient or Capital Cost <strong>of</strong> a component.<br />

Greeks<br />

α<br />

β<br />

γ<br />

δ<br />

∆<br />

ε<br />

η<br />

κ<br />

λ<br />

µ<br />

ν<br />

π<br />

ρ<br />

φ<br />

ℵ<br />

ϕ<br />

ϖ<br />

Sonic velocity or constant <strong>of</strong> Cobb-Douglass equation.<br />

Constant for calculating vapor entropy.<br />

Activity coefficient.<br />

Interstage (temperature) difference.<br />

Difference, increment, variation (or loss).<br />

Relative error or ratio.<br />

Efficiency.<br />

Technical production coefficient.<br />

Latent heat, real number or Lagrange multiplier.<br />

Viscosity or chemical exergy component.<br />

Specific volume.<br />

Osmotic pressure.<br />

Density.<br />

Arrays/Matrices<br />

B<br />

[DF]<br />

DF<br />

DI<br />

∆FT<br />

Mass flow coefficient <strong>of</strong> a turbine section, or dysfunction coefficient.<br />

Constant for calculating latent heat <strong>of</strong> vapor.<br />

Amortization factor.<br />

Chamber load or total final product.<br />

Exergy flows set.<br />

Dysfunction matrix.<br />

Array <strong>of</strong> dysfunctions generated in the components.<br />

Array <strong>of</strong> dysfunctions generated by the components.<br />

Impact on fuel array.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 437


438<br />

Nomenclature<br />

κe<br />

KD<br />

〈 KP〉<br />

MF<br />

I<br />

| I〉<br />

P<br />

P<br />

S<br />

| P〉<br />

U<br />

D<br />

Subscripts<br />

Unit exergy consumption array <strong>of</strong> the system input resources.<br />

Diagonal matrix <strong>of</strong> the unit exergy consumption.<br />

Unit exergy consumption matrix.<br />

Malfunction array.<br />

Irreversibility array.<br />

Irreversibility matrix operator.<br />

Product array.<br />

Final product array.<br />

Product matrix operator.<br />

Unitary matrix.<br />

a Absolute.<br />

b Exergy flow or brine.<br />

B Brine.<br />

bi Brine inside the tubes.<br />

c Condensate.<br />

C Condenser.<br />

ci Steam to Ejector from leakage system.<br />

CT Condensing Turbine.<br />

d Distillate, design.<br />

D Distillate.<br />

des Low-Pressure Steam to MSF unit.<br />

DR Deaerator.<br />

e Exit or electricity.<br />

es Interstage.<br />

ex Extraction.<br />

f Fouling, formation or fuel relative.<br />

F Cooling brine.<br />

fg Evaporation.<br />

fm Film.<br />

gen Generator.<br />

H Brine Heater.<br />

H2O<br />

Pure water.<br />

i Inlet, i-section or array index.<br />

j j-Stage, index, variable or specie.<br />

K Kelvin.<br />

L Loss.<br />

ls Live Steam Flow.<br />

LS Live Steam Extraction from reduction pressure station.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Nomenclature<br />

lm Logaritmic mean.<br />

m Mean.<br />

msf MSF plant.<br />

N Last stage <strong>of</strong> MSF unit.<br />

NRC Last stage <strong>of</strong> recovery section.<br />

o Outlet.<br />

P Demister pressure losses, or product.<br />

q Chemical.<br />

r Reference.<br />

rcs Recovery section (exit).<br />

rdes Condensate returned from the MSF unit (heater), after passing brine heater<br />

pump.<br />

s Isoentropic, shell or entropy flow.<br />

S Saturated.<br />

sea Seawater.<br />

ST Steam or Steam Turbine.<br />

t Turbine or tube.<br />

T Total.<br />

va Steam to vacuum system <strong>of</strong> MSF unit (condensate returning to condenser).<br />

vent Venting system.<br />

w Wall or water.<br />

Z Capital cost.<br />

0 To the environment.<br />

Superscripts<br />

a, b, c, x, y, z Exponents for calculations <strong>of</strong> TTDs in heaters or deaerator, pressure losses or<br />

gl<strong>and</strong> steam system.<br />

L Local.<br />

G Induced.<br />

m m-Iteration or scaling factor.<br />

n1, n2, n3, n4 Exponents for capital costing equation.<br />

´ Extraction mass flow rate.<br />

o St<strong>and</strong>ard state.<br />

r Operating parameter.<br />

t Transpose (matrix notation).<br />

–1 Inverse (matrix notation).<br />

0 Reference or design (matrix notation).<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 439


References<br />

Asea Brown Bovery (ABB) (1996a).<br />

Details. Private Communication.<br />

Boiler Performance Data. Construction<br />

Asea Brown Bovery (ABB) (1996b). Diagram Charts, <strong>and</strong> Heat Balance <strong>of</strong><br />

Al Taweelah B Power Generation Unit. Private Communication.<br />

Asea Brown Bovery (ABB) (1996c). HP <strong>and</strong> LP Heaters, Feedwater Storage Tank,<br />

Cold Storage Tank. Private Communication.<br />

Asea Brown Bovery (ABB) (1996d). Steam, Condensate <strong>and</strong> Feedwater Piping<br />

System. Private Communication.<br />

Asea Brown Bovery (ABB) (1996e). Generator System. Private Communication.<br />

Asea Brown Bovery (ABB) (1996f). Pump Curves. Private Communication.<br />

Asea Brown Bovery (ABB) (1997). Private communication.<br />

Abdel-Jawad, M., Al-Tabtabaei, M. (1999). Impact on Current Power Generation<br />

<strong>and</strong> Water Desalination Activities on Kuwait Marine Environment. Proceedings <strong>of</strong><br />

the IDA World Congress on Desalination <strong>and</strong> Water Reuse. San Diego, USA.<br />

Abu Qdais, H. A. (1999). Environmental Impacts <strong>of</strong> Desalination Plants on the<br />

Arabian Gulf.<br />

Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong> Water<br />

Reuse. San Diego, USA.<br />

Afgan, N. H., Darwish, M., Carvalho, M. G. (1999). Sustainability Assessment <strong>of</strong><br />

Desalination Plants for Water Production. Desalination 124, pp. 19-32. Presented at<br />

the European Conference on Desalination <strong>and</strong> the Environment. Las Palmas, Spain.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


442<br />

References<br />

Alawadhi, A. A. (1999). Regional Report on Desalination. Proceedings <strong>of</strong> the IDA<br />

World Congress on Desalination <strong>and</strong> Water Reuse. San Diego, USA.<br />

Al-Gobaisi, D. M. K. (1999). Water for Sustainable Development <strong>of</strong> the Arab World.<br />

Private Communication.<br />

Al-Gobaisi, D. M. K. (1997). Sustainable Augmentation <strong>of</strong> Fresh Water Resources<br />

through Appropriate Energy <strong>and</strong> Desalination Technologies. Proceedings <strong>of</strong> the IDA<br />

World Congress on Desalination <strong>and</strong> Water Reuse. Madrid, Spain.<br />

Alhumaizi, K. (1997). Modeling, Simulation <strong>and</strong> Control <strong>of</strong> Multistage Flash<br />

Desalination Plant,<br />

Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong><br />

Water Reuse. Madrid, Spain. Vol. III, pp. 111-130.<br />

Al-Mutaz, I. S., Soliman, M. A. (1989). Simulation <strong>of</strong> MSF Desalination Plants.<br />

Desalination 74, pp. 317-326.<br />

Al-Owais, A. A., Nijhawan, R. K., Budhiraja, P. K. (1989). Operational Experience<br />

<strong>of</strong> Once Through MSF Desalination Units. Desalination 73, pp. 327-340.<br />

Al-Sulaiman, F. A., Ismail, B. (1995). Exergy Analysis <strong>of</strong> Major Recirculating<br />

Multi-stage Flash Desalting Plants in Saudi Arabia. Desalination 103, pp. 265-270.<br />

3<br />

Andrews, T., Shumway, S. A. (1999). Design Study <strong>of</strong> a 20,000 m /day Seawater<br />

Reverse Osmosis Work Exchanger Energy Recovery System. Proceedings <strong>of</strong> the IDA<br />

World Congress on Desalination <strong>and</strong> Water Reuse. San Diego, USA.<br />

Ahrendts, J. (1980). Reference States. Energy 5, Vol. 8, pp. 667-677.<br />

American Society <strong>of</strong> Mechanical Engineers (ASME) (1967). 1967 ASME Steam<br />

Tables. The American Society <strong>of</strong> Mechanical Engineers. New York, USA.<br />

Badr, O., Probert, S. D., O’Callaghan, P. (1990). Rankyne Cycles for Steam Power-<br />

Plants. Applied Energy 36, pp. 191-231.<br />

Barba, D., Liuzzo, G., Tagliaferri, G. (1973). Mathematical Model for Multiflash<br />

th<br />

Desalting Plant Control. 4 International Symposium on Fresh Water from the Sea.<br />

Vol. 1, pp. 153-168.<br />

Barendsen, W. C., Moch, I. (1999). Privatization <strong>of</strong> Seawater Reverse Osmosis<br />

Plants in Antigua. Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong><br />

Water Reuse. San Diego, USA.<br />

Barner, H. E., Scheuerman, R. V. (1978). H<strong>and</strong>book <strong>of</strong> Thermochemical Data for<br />

Compounds <strong>and</strong> Acqueous Species. John Wiley <strong>and</strong> Sons Inc.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


References<br />

Barthelmes, J., Bolmer, H. (1996). Fouling <strong>and</strong> Scaling Control in MSF<br />

Desalination Units by “On-Load” Tube Cleaning. Desalination <strong>and</strong> Water Reuse,<br />

Vol. 7/2, pp. 27-33.<br />

Brown Boveri Co. (BBC) (1979). Thermal Kit, Simplified Instructions for the<br />

Thermal Calculation <strong>of</strong> the Antikokan Units.<br />

HTGD 12246 E.<br />

Beamer, J. H., Wilde, J. D. (1971). The Simulation <strong>and</strong> Optimization <strong>of</strong> a Single<br />

Effect Multi-Stage Flash Desalination Plant.<br />

Desalination 9, pp. 259-275.<br />

Bejan, A., Tsatsaronis, G., Moran, M. (1997). Thermal Design <strong>and</strong> Optimization.<br />

John Wiley <strong>and</strong> Sons Inc., New York.<br />

Benelmir, R. (1989). Second Law Analysis <strong>of</strong> a Co-generation Cycle. Ph. D. Thesis.<br />

Georgia Institute <strong>of</strong> Technology.<br />

Boehm, R. F. (1987). Design Analysis <strong>of</strong> Thermal Systems. Ed. John Wiley <strong>and</strong><br />

Sons. New York.<br />

Br<strong>and</strong>ani, V., Del Re, G., Di Giacomo, G. (1985). A New Model for Predicting<br />

Thermodynamic Properties <strong>of</strong> Sea Salt Solutions. Desalination 56, pp. 299-313.<br />

Breidenbach, L., Rautenbach, R., Tusel, G. F. (1997). <strong>Thermoeconomic</strong> Assessment<br />

<strong>of</strong> Fossil Fuel Fired Dual Purpose Power/Water Plants.<br />

Proceedings <strong>of</strong> the IDA<br />

World Congress on Desalination <strong>and</strong> Water Reuse. Madrid, Spain. Vol. IV,<br />

pp. 167-180.<br />

Bromley, L. A., Diamond, A. E., Salam, E., Wilkins, D. G. (1970). J. Chem. Eng.<br />

Data 15, pp. 246.<br />

Brodyansky, V., B<strong>and</strong>ura, A. (1993). The Prognosis for Macroeconomical<br />

Development <strong>and</strong> Exergy.<br />

Proceedings <strong>of</strong> the International Symposium on energy<br />

systems <strong>and</strong> ecology ENSEC’93. Cracow, Pol<strong>and</strong>. pp. 153-161.<br />

Cadagua (1999). Private Communication.<br />

Calder (1999). Pelton Wheel Energy Recovery Turbines. Private Communication.<br />

Chen, S. F., Chan, R. C., Read, S. M., Bromley, L. A. (1973). Viscosity <strong>of</strong> Sea Water<br />

Solutions. Desalination 13, pp. 37-51.<br />

Coleman, A. K. (1971). Optimization <strong>of</strong> a Single Effect, Multi-Stage Flash<br />

Distillation Desalination System. Desalination 9, no. 4, pp. 315-331.<br />

Cooke, D. H. (1985). On prediction <strong>of</strong> Off-Design Multistage Turbine Pressures by<br />

Stodola’s Ellipse. Journal <strong>of</strong> Engineering for Gas Turbines <strong>and</strong> Power. Transactions<br />

<strong>of</strong> the ASME. Vol. 107, pp. 596-606.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

443


444<br />

References<br />

Corripio, A. B., Chrien, K. S., Evans, L, B. (1982). Estimate Costs <strong>of</strong> Heat<br />

Exchangers <strong>and</strong> Storage Tanks Via Correlations. Chemical Engineering, January<br />

1982, pp. 125-127.<br />

Cotton, K. C. (1993). Evaluating <strong>and</strong> Improving Steam Turbines Performance,<br />

Edit.<br />

Cotton Factory Inc. New York, USA.<br />

Darwish, M. A., Al-Najem, N. M., Al-Ahmad, M. S. (1993). Second-Law Analysis <strong>of</strong><br />

Recirculating Multi-stage Flash System. Desalination 89, pp. 289-309.<br />

Darwish, M. A., Yousef, F. A., Al-Najem, N. M. (1997). Energy Consumption <strong>and</strong><br />

Costs with a Multi-stage Flashing (MSF) Desalting System. Desalination 109,<br />

pp. 285-302.<br />

Darwish, M. K., Arazzini, S. (1989). Description <strong>and</strong> Mathematical Model <strong>of</strong> a<br />

Large MSF Desalination Plant in Scada Configuration.<br />

Private Communication,<br />

pp. 91-106.<br />

De Armas, J. C., Pérez, J. L., Von Gottberg, A. J. M. (1999). Desalination <strong>of</strong><br />

Municipal Sewage Effluent with Electrodialysis Reversal in Tenerife.<br />

Proceedings <strong>of</strong><br />

the IDA World Congress on Desalination <strong>and</strong> Water Reuse. San Diego, USA.<br />

Echaniz, J., Rodero, A., Sallangos, O., Santamaria F. J. (1997). Dhekelia (Cyprus)<br />

Seawater Desalination Plant Design, Construction <strong>and</strong> Commissioning <strong>of</strong> the<br />

3 20,000 m /day R.O. Plant. Proceedings <strong>of</strong> the IDA World Congress on Desalination<br />

<strong>and</strong> Water Reuse. Madrid, Spain. Vol. II, pp. 371-392.<br />

El-Nashar, A. M. (1999). Cost Allocation in a Cogeneration Plant for the<br />

Production <strong>of</strong> Power <strong>and</strong> Desalted Water – Comparison <strong>of</strong> the Exergy Cost<br />

Accounting Method with the WEA Method.<br />

Private Communication.<br />

El-Nashar, A. M., Qamhiyeh, A. A. (1993). Optimal Performance <strong>of</strong> MSF Distillers<br />

for UANW 9 <strong>and</strong> 10 Power Plant: A <strong>Thermoeconomic</strong> Study.<br />

Desalination 93,<br />

pp. 323-342.<br />

Elovic, P., Willocks, G. (1999). Case Study <strong>of</strong> Operating Experience <strong>of</strong> 9 Low<br />

Temperature MED plants in the U.S. Virgin Isl<strong>and</strong>s. Proceedings <strong>of</strong> the IDA World<br />

Congress on Desalination <strong>and</strong> Water Reuse. San Diego, USA.<br />

El-Saie, M. H., El-Saie, Y. M. H. (1989). Optimization <strong>of</strong> Dual-Purpose Steam<br />

Power <strong>and</strong> MSF Desalination Plant. Desalination 76, pp. 155-175.<br />

El-Sayed, Y. M., Aplenc, A. J. (1970). Application <strong>of</strong> the <strong>Thermoeconomic</strong> Approach<br />

in the Analysis <strong>and</strong> Optimization <strong>of</strong> Vapor-Compression Desalting System.<br />

Transactions <strong>of</strong> the ASME. Journal <strong>of</strong> Engineering <strong>and</strong> Power, Vol. 92 no. 1,<br />

pp. 17-26.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


References<br />

El-Sayed, Y. M., Evans, R. B. (1970). <strong>Thermoeconomic</strong>s <strong>and</strong> the Design <strong>of</strong> Heat<br />

Systems.<br />

Transactions <strong>of</strong> the ASME. Journal <strong>of</strong> Engineering <strong>and</strong> Power, Vol. 92<br />

no. 1, pp. 27-35.<br />

El-Sayed, Y. M., Silver, R. S. (1980). Fundamentals <strong>of</strong> Distillation.<br />

Principles <strong>of</strong><br />

Desalination, Chapter 2. Academic Press Inc.<br />

El-Sayed, Y. M., Tribus, M. (1983). Strategic use <strong>of</strong> <strong>Thermoeconomic</strong>s for Systems<br />

Improvement. ACS Symposium Series, no. 235, pp. 215-238. Washington D.C.,<br />

USA.<br />

El-Sayed, Y. M. (1988). A Decomposition Strategy for <strong>Thermoeconomic</strong>s<br />

Optimization <strong>of</strong> a Given New Configuration,<br />

Approaches to the Design <strong>and</strong><br />

Optimization <strong>of</strong> Thermal Systems. Wepfer <strong>and</strong> Moran eds. ASME, pp. 41-47. New<br />

York, USA.<br />

El-Sayed, Y. M. (1996). Second-Law-Based Analysis <strong>and</strong> Optimization <strong>of</strong> Seawater<br />

Desalting Systems. Private Communication.<br />

Erbes, M. R., Gay, R. B. (1989). Gate/cycle Predictions <strong>of</strong> the Off-Design<br />

Performance <strong>of</strong> Combined Cycle Power Plants. ENTER S<strong>of</strong>tware, Inc. ASME 1989<br />

WAM.<br />

Erlach, B. (1998). Comparison <strong>of</strong> <strong>Thermoeconomic</strong> Methodologies: Structural<br />

Theory, AVCO <strong>and</strong> LIFO. Application to a Combined Cycle. University <strong>of</strong> Zaragoza.<br />

Dept. <strong>of</strong> Mechanical Engineering.<br />

Erlach, B., Serra, L., Valero, A. (1999). Structural Theory as St<strong>and</strong>ard for<br />

<strong>Thermoeconomic</strong>s. Energy Conversion <strong>and</strong> Management 40, pp. 1627-1649.<br />

Ettouney, H. M., El-Dessouky, H. T. (1999). A Simulator for Thermal Desalination<br />

Processes. Desalination 125, pp. 277-292. Presented at the European Conference on<br />

Desalination <strong>and</strong> the Environment. Las Palmas, Spain.<br />

Evans, R. B. (1962). A Contribution to the Theory <strong>of</strong> Thermo-Economics. Sea Water<br />

Research Project S. W. 604. Report no. 62-36. Department <strong>of</strong> Engineering.<br />

University <strong>of</strong> California.<br />

Evans, R. B., Crellin, G. L., Tribus, M. (1980). <strong>Thermoeconomic</strong> Considerations <strong>of</strong><br />

Sea Water Demineralization.<br />

Principles <strong>of</strong> Desalination, Chapter 1. Academic Press<br />

Inc.<br />

Evans, R. B. (1980). <strong>Thermoeconomic</strong> Isolation <strong>and</strong> Essergy Analysis.<br />

Energy,<br />

Vol. 5, no. 8-9, pp. 805-822.<br />

Fabuss, B. M., Korosi, A. (1968). Properties <strong>of</strong> Sea Water <strong>and</strong> Solutions Containing<br />

Sodium Chloride, Potassium Chloride, Sodium Sulfate <strong>and</strong> Magnesium Sulfate.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

445


446<br />

References<br />

Research <strong>and</strong> Development Progress Report no. 384. U.S. Department <strong>of</strong> the<br />

Interior.<br />

Falceta, F., Sciubba, E. (1997). Modelling <strong>and</strong> Simulation <strong>of</strong> Multi-Stage Flash<br />

Desalination Plants.<br />

Private Report.<br />

Fayas, J., Novoa, J. (1997). The Desalination Process in the Balearic Isl<strong>and</strong>s.<br />

Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong> Water Reuse. Madrid,<br />

Spain. Vol I, pp. 41-54.<br />

Fisia Italimpianti (1996). Private Communication.<br />

Fisia Italimpianti (1997).<br />

Communication.<br />

Diagram charts <strong>of</strong> the MSF Plant.<br />

Fisia-Italimpianti (1999). Water Desalination Plants. Private Communication.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

Private<br />

Frangopoulos, C. (1983). <strong>Thermoeconomic</strong> Functional Analysis: A method for the<br />

Optimal Design or Improvement <strong>of</strong> Complex Thermal Systems. Ph. D. Thesis.<br />

Georgia Institute <strong>of</strong> Technology.<br />

Frangopoulos, C. A. (1987). <strong>Thermoeconomic</strong> Functional Analysis <strong>and</strong><br />

Optimization. Energy Vol. 12, no. 7, pp. 563-571.<br />

Frangopoulos, C. A. (1988). Optimal Design <strong>of</strong> a Gas Turbine Plant by a<br />

<strong>Thermoeconomic</strong> Approach. ASME COGEN-TURBO, pp. 563-571. ASME, New<br />

York, USA.<br />

Frangopoulos, C. A. (1990). Intelligent Functional Approach: A Method for Analysis<br />

<strong>and</strong> Optimal Synthesis-Design-Operation <strong>of</strong> Complex Systems. Proceedings <strong>of</strong> the<br />

International Symposium: A future for energy. Florence, Italy. Pergamon Press,<br />

pp. 805-815.<br />

Frangopoulos, C. A. (1991). Private Communication.<br />

Frangopoulos, C. A. (1994). Application <strong>of</strong> the <strong>Thermoeconomic</strong> Functional<br />

Approach to the CGAM Problem.<br />

Energy Vol. 19, no. 13.<br />

Friedrich, R. O., Hafford, J. A. (1971). Report ORNL-TM-3489.<br />

Gaggioli, R. A. (1980). Thermodynamics: Second Law Analysis.<br />

ACS Symposium<br />

Series 122. American Chemical Society. Washington D.C., USA.<br />

Gaggioli, R. A., El-Sayed, Y. M. (1987). A Critical Review <strong>of</strong> Second Law Costing<br />

Methods.<br />

Proceedings <strong>of</strong> the IV International Symposium on Second Law Analysis<br />

<strong>of</strong> Thermal Systems (ASME Book I00236). ASME, pp. 59-73. New York, USA.


References<br />

García, L., Gómez, C. (1999). Conditions for Economical Benefits <strong>of</strong> the Use <strong>of</strong><br />

Solar Energy in Multi-stage Flash Distillation. Desalination 125, pp. 133-138.<br />

Presented at the European Conference on Desalination <strong>and</strong> the Environment. Las<br />

Palmas, Spain.<br />

García, L., Palmero, A. I., Gómez, C. (1999). Application <strong>of</strong> Direct Steam<br />

Generation into a Solar Parabolic Trough Collector to Multieffect Distillation.<br />

Desalination 125, pp. 139-145. Presented at the European Conference on<br />

Desalination <strong>and</strong> the Environment. Las Palmas, Spain.<br />

Gleick, P. H. (1998). The World’s Water – The Biennial Report on Freshwater<br />

Resources, 1998/1999.<br />

Isl<strong>and</strong> Press. Washington DC, USA.<br />

Glueck, A. R., Bradshaw, R. W. (1970). A Mathematical Model for a Multistage<br />

rd<br />

Flash Distillation Plant. 3 International Symposium on Fresh Water from the Sea.<br />

Vol. 1, pp. 95-108.<br />

Georgescu-Roegen, N. (1971). The Entropy Law <strong>and</strong> the Economic Process.<br />

Harvard University Press. Cambridge MA. USA.<br />

Goto, T., MacCormick, T., Congjie, G., Guoling, R., Chung Y.-T.<br />

(1999). Overview<br />

<strong>of</strong> Desalination in the Pacific Region.<br />

Proceedings <strong>of</strong> the IDA World Congress on<br />

Desalination <strong>and</strong> Water Reuse. San Diego, USA.<br />

Hamed, O. A., Al-S<strong>of</strong>i, M. A. K., Iman, M., Mustafa, G. M., Ba-Mardouf, K., Al-<br />

Washmi, H. (1999). Thermal Performance <strong>of</strong> Multistage Flash Distillation Plants in<br />

Saudi Arabia. Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong> Water<br />

Reuse. San Diego, USA.<br />

Hanbury, W. T., Hodgkiess, T., Morris, R. (1993). Desalination Technology 93. An<br />

Intensive Course.<br />

Porthan Ltd., Easter Auchinloch. Lenzie, Glasgow, UK.<br />

Hassan, A. S., Florido, P. C. (1999). Feasibility <strong>of</strong> Nuclear Desalination Costs in<br />

Egypt. Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong> Water Reuse.<br />

San Diego, USA.<br />

Hauge, L. J., Ludvigsen, F. (1999). Field Installation <strong>of</strong> Pressure Exchanger in a<br />

3 80 m /d SWRO Plant. Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong><br />

Water Reuse. San Diego, USA.<br />

Hayakawa, K., Satori, H., Konishi, K. (1973). Process Simulation on a Multi-Stage<br />

Flash Distillation Plant. 4 th International Symposium on Fresh Water from the Sea.<br />

Vol. 1, pp. 303-312.<br />

Helal, A. M., Medani, M. S., Soliman, M. A. (1986). A Tridiagonal Matrix Model<br />

for MultiStage Flash Desalination Plants. Computers & Chemical Engineering Vol.<br />

10, pp. 95-108.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

447


References<br />

Hömig, H. E. (1978). Seawater <strong>and</strong> Seawater Distillation (Fichtner H<strong>and</strong>book).<br />

Vulkan-Verlag. Essen, Germany.<br />

Husain, A., Hassan, A., Al-Gobaisi, D. M. K., Al-Radif, A., Woldai, A., Sommariva,<br />

C. (1993). Modelling, Simulation, Optimization <strong>and</strong> Control <strong>of</strong> Multistage Flashing<br />

(MSF) Desalination Plants. Part I: Modelling <strong>and</strong> Simulation. Desalination 92,<br />

pp. 21-41.<br />

Husain, A., Woldai, A., Al-Radif, A., Kesou, A., Borsani, R., Sultan, H.,<br />

Deshp<strong>and</strong>ey, P. B. (1994). Modelling, <strong>and</strong> Simulation <strong>of</strong> a Multistage Flash (MSF)<br />

Desalination Plant. Desalination 97, pp. 555-586.<br />

Husain, A. (1999). Computational Aspects in Process Simulation. International<br />

Center for Water <strong>and</strong> Energy Systems (ICWES). Private Communication.<br />

I.D.E. Technologies ltd. (1999). Private Communication.<br />

Intermón (1998). Relaciones Norte-Sur. Conceptos Clave. Dossiers para entender el<br />

mundo. Ed. Octaedro.<br />

Isdale, J. D., Spence, C. M., Tudhope, J. S. (1972). Physical Properties <strong>of</strong> Sea Water<br />

Solutions: Viscosity. Desalination 10, pp. 319-328.<br />

Itahara, S., Stiel, L. Y. (1968). The Optimal Design <strong>of</strong> Multi-Stage Flash<br />

Evaporators by Dynamic programming. Desalination 4, pp. 248-257.<br />

Jernqvist, A., Jernqvist, M., Aly, G. (1999). Simulation <strong>of</strong> Thermal Desalination<br />

Processes. Desalination 126, pp. 147-152. Presented at the European Conference on<br />

Desalination <strong>and</strong> the Environment. Las Palmas, Spain.<br />

JSME (1968). 1968 JSME Steam Tables. The Japan Society <strong>of</strong> Mechanical<br />

Engineers. Tokyo, Japan.<br />

Keenan, J. H., Keyes, F. G. (1955). Thermodynamic Properties <strong>of</strong> Steam. John Wiley<br />

<strong>and</strong> Sons Inc., New York, USA.<br />

Keenan, J. H., Keyes, F. G., Hill, F., Moore, K. (1955). Steam Tables (SI Units). John<br />

Wiley <strong>and</strong> Sons Inc., New York, USA.<br />

Klein, S. A., Alvarado, F. L. (1992). EES Engineering Equation Solver. F-Chart<br />

S<strong>of</strong>tware. Middleton, USA.<br />

Kotas, T. (1985). The Exergy Method <strong>of</strong> Thermal Plant Analysis. Butteerworth eds.,<br />

London, UK.<br />

Kronenberg, G., Dvornikov, V. (1999). Fuel Cost <strong>of</strong> Water (FCW) in Dual Plants.<br />

Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong> Water Reuse. San<br />

Diego, USA.<br />

448 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


References<br />

Lazzaretto, A., Tsatsaronis, G. (1997). On the Quest for Objective Equations in<br />

Exergy Costing. Proceedings <strong>of</strong> the ASME Advanced Energy Systems Division.<br />

ASME. AES-Vol. 37, pp. 197-210.<br />

Le G<strong>of</strong>f, P. (1979). Energetique Industrièlle. Tecnique et Documentation. Paris,<br />

France.<br />

Lerch, F., Royo, J., Serra, L. (1999). Structural Theory <strong>and</strong> <strong>Thermoeconomic</strong><br />

Diagnosis. Part II: Application to an Actual Power Plant. Proceedings <strong>of</strong> the<br />

ECOS’99 Conference. ASME, pp. 374-379. Tokyo, Japan.<br />

Lewis, G. N., R<strong>and</strong>al, M. (1961). Thermodynamics. Mc-Graw Hill Company.<br />

Leyendekkers, J. V. (1979). Prediction <strong>of</strong> the Density <strong>and</strong> Viscosity <strong>of</strong> Seawater its<br />

Concentrates <strong>and</strong> other Multicomponents Solutions Using the Tamman-Tait-Gibson<br />

(TTG) Model. Desalination 29, pp. 263-274.<br />

Lozano, M. A., Valero, A. (1993). Theory <strong>of</strong> the Exergetic Cost. Energy, Vol. 18,<br />

no. 3, pp. 939-960. Elsevier Science Ltd., UK.<br />

Lozano, M. A., Valero, A., Serra, L. (1993). Theory <strong>of</strong> the Exergetic Cost <strong>and</strong><br />

<strong>Thermoeconomic</strong>s Optimization. Proceedings <strong>of</strong> the International Symposium<br />

ENSEC’93. Cracow, Poll<strong>and</strong>.<br />

Lozano, M. A., Bartolomé, J. L., Valero, A., Reini, M. (1994). <strong>Thermoeconomic</strong><br />

Diagnosis <strong>of</strong> Energy Systems. Flowers 94, Florence World Energy Research<br />

Symposium, pp. 149-156. Florence, Italy.<br />

Lozano, M. A., Valero, A., Serra, L. (1996). Local Optimization <strong>of</strong> Energy Systems.<br />

Proceedings <strong>of</strong> the ASME Advanced Energy System Division. Atlanta, Georgia.<br />

AES-Vol. 36, pp. 241-250.<br />

Marín, J. M., Turégano, J. A. (1985). Contribution to the Calculation <strong>of</strong> Chemical<br />

Exergy in Industrial Processes (Electrolyte Solutions). Energy Vol. 11, pp. 231-236.<br />

Martin, M. H. (1919). Articles on Leakage <strong>of</strong> Steam Through Dummy Pistons.<br />

Engineering Jan 3.<br />

Martínez, A., Serra, L., Valero, A. (2000). Cost Assessing in Entrained Flow<br />

Gasifiers Based on Physical Models. Paper accepted to ECOS’2000 Conference.<br />

Enschede, Netherl<strong>and</strong>s.<br />

Menéndez, E. (1997). Las Energías Renovables. Los Libros de la Catarata Eds.<br />

Medina, J. A. (2000). Desalación de Aguas Salobres y de Mar. Osmosis Inversa.<br />

Mundi-Prensa Eds. Madrid, Spain.<br />

Micros<strong>of</strong>t Corporation (1997). Micros<strong>of</strong>t Fortran Developer Studio.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 449


References<br />

Micros<strong>of</strong>t Corporation <strong>and</strong> Cooper S<strong>of</strong>tware Inc. (1997). Micros<strong>of</strong>t Visual Basic<br />

Version 5.0.<br />

Moran, M. J. (1990). Second Law Analysis. What is the State <strong>of</strong> the Art?<br />

International Symposium: A future for energy. Florence (Italy). Pergamon Press,<br />

pp. 249-260.<br />

Morris, D. R., Szargut, J. (1986). St<strong>and</strong>ard Chemical Exergy <strong>of</strong> some Elements <strong>and</strong><br />

Compounds on the Planet Earth. Energy Vol. 11, pp. 733-749.<br />

Mothersed, C. T. (1966). Report ORNL-TM-1560, August 1966.<br />

National Office <strong>of</strong> Potable Water (NPOW) (1996). Non Conventional Waters Use for<br />

Drinking Water Supply/Seawater Desalination. Kingdom <strong>of</strong> Morocco. Private<br />

Communication.<br />

Newman, A. (1980). Thermodynamics <strong>of</strong> Acqueous Systems with Industrial<br />

Applications. American Chemical Society. Washington DC, USA.<br />

Omar, A. M. (1981). M. Sc. Thesis. UPM Dharan (Saudi Arabia).<br />

Ophir, A., Gendel, A. (1999). Development <strong>of</strong> the World’s Largest Multi-Effect<br />

Mechanical Vapor Compression (M.E.M.V.C.) Desalination Plants. Proceedings <strong>of</strong><br />

the IDA World Congress on Desalination <strong>and</strong> Water Reuse. San Diego, USA.<br />

Perry, R. H., Chilton, C. (1984). Manual del Ingeniero Químico. Ed. McGraw-Hill,<br />

5 th edition. Vol. I, pp. 446.<br />

Pina, H. L. G. (1979). A Computer Program for the Calculation <strong>of</strong> the<br />

Thermodynamic Properties <strong>of</strong> Water. Revue Generale de Thermique, no. 215,<br />

pp. 689-693.<br />

Pisa, J. (1997). <strong>Thermoeconomic</strong> Analysis <strong>of</strong> IGCC plants. Ph. D. Thesis.<br />

Department <strong>of</strong> Mechanical Engineering. University <strong>of</strong> Zaragoza.<br />

Ponce, S. L., Jankel, L. H. (1999). The Value <strong>of</strong> Water in the 21 st Century – Impact<br />

on U.S. Desalination. Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong><br />

Water Reuse. San Diego, USA.<br />

Powell, M. J. D. (1964). An Efficient Method for Finding the Minimun <strong>of</strong> a Function<br />

<strong>of</strong> Several Variables without Calculate Derivatives. Computer J. Vol. 7, pp. 155-162.<br />

Prabhakar, S., Hanra, M. S., Misra, B. M., Sadhukan, H. K. (1997). Small<br />

Desalination Units to Provide Safe Drinking Water in Remote Rural Areas in India.<br />

Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong> Water Reuse. Madrid,<br />

Spain. Vol. I, pp. 3-16.<br />

450 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


References<br />

Rautenbach, R., Buchel, H. G. (1979). Modular Program for Design <strong>and</strong> Simulation<br />

<strong>of</strong> Desalination Plants. Desalination 31, pp. 71-83.<br />

Reid, R. C., Prausnitz, J. M., Sherwood, T. K. (1977). The Properties <strong>of</strong> Gases <strong>and</strong><br />

Liquids. McGraw-Hill. New York, USA.<br />

Reini, M. (1994). Analisi e Sviluppo dei Metodi Termoeconomi per lo Studio degli<br />

Impianti di Conversione dell’Energia. Ph. D. Thesis. Università di Padova.<br />

Royo, F. J. (1994). Las Ecuaciones Características de los Sistemas Térmicos. La<br />

energía Libre Relativa. Ph. D. Thesis. Dept. <strong>of</strong> Mechanical Engineering. University<br />

<strong>of</strong> Zaragoza.<br />

Saeed, M. N. (1992). Fuel Efficiencies, Allocation <strong>of</strong> Fuels <strong>and</strong> Fuel Cost for Power<br />

<strong>and</strong> Desalination in Dual Purpose Plants: A Novel Methodology. Desalination 85,<br />

pp. 213-229.<br />

Salisbury, J. K. (1974). Steam Turbines <strong>and</strong> their Cycles. Krieger Publishing<br />

Company. New York, USA.<br />

Sánchez, J. M., Velasco, J., Kindelan, J. M., Andreu, J. (1997). Marbella Seawater<br />

Desalination Plant: Construction <strong>and</strong> Start-up Experience. Proceedings <strong>of</strong> the IDA<br />

World Congress on Desalination <strong>and</strong> Water Reuse. Madrid, Spain. Vol. V,<br />

pp. 463-478.<br />

Schnakel, H. C. (1958). Formulations for the Thermodynamic Properties <strong>of</strong> Steam<br />

<strong>and</strong> Water. Trans. ASME 80, pp. 959-966.<br />

Sengers, J. V., Watson, J. T. R. (1986). Improved International Formulations for the<br />

Viscosity <strong>and</strong> Thermal Conductivity <strong>of</strong> Water Substance. Journal <strong>of</strong> Physical &<br />

Chemical Reference Data, Vol. 15, no. 4, pp. 1291-1314.<br />

Sephton, H. H., Solomon, R. L. (1997). Use <strong>of</strong> Power Plant Turbine Reject Steam to<br />

Drive Desalination with Enhanced Heat Transfer Performance. Proceedings <strong>of</strong> the<br />

IDA World Congress on Desalination <strong>and</strong> Water Reuse. Madrid, Spain. Vol. IV, pp.<br />

299-308.<br />

Sephton, H. H. (1999). Turbine Exhaust Steam Driven Desalination. Proceedings <strong>of</strong><br />

the IDA World Congress on Desalination <strong>and</strong> Water Reuse. San Diego, USA.<br />

Serra, L. (1994). Optimización Exergoeconómica de Sistemas Térmicos.<br />

Ph. D. Thesis. Department <strong>of</strong> Mechanical Engineering. University <strong>of</strong> Zaragoza.<br />

Slesarenko, V. N., Shtim, A. S. (1987). Exergy Analysis <strong>of</strong> Multi-Stage Desalination<br />

Plants. Desalination 61, pp. 1-5.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 451


References<br />

Slesarenko, V. N. (1999). Desalination Plant with Absorption Heat Pump for Power<br />

Station. Desalination 126, pp. 281-286. Presented at the European Conference on<br />

Desalination <strong>and</strong> the Environment. Las Palmas, Spain.<br />

Spanish Desalination <strong>and</strong> Water Reuse Association (AEDyR) (1999). Private<br />

Communication.<br />

Spencer, R. C., Cotton, K. C., Cannon, C. N. (1974). A Method <strong>of</strong> Predicting the<br />

Performance <strong>of</strong> Steam Turbine Generators 16.500 kW <strong>and</strong> Larger. General Electric<br />

Corp., Publication GER-2007C.<br />

Spiegler, K. S., El-Sayed, Y. M. (1994). A Desalination Primer. Balaban<br />

Desalination Publications. Italy.<br />

Stodola, A. (1927). Steam <strong>and</strong> Gas Turbines. McGraw Hill Company, Vol. 1,<br />

pp. 316. New York, USA.<br />

Stoughton, R. W., Lietzke, M. H. (1965). J. chem. Engng. Data 10, pp. 254.<br />

Szargut, J. (1980). International Progress in Second Law Analysis. Energy Vol. 5,<br />

pp. 709-718.<br />

Szargut, J. (1989). Chemical Exergies <strong>of</strong> the Elements. Applied Energy 32, pp. 269-<br />

286.<br />

Tadros, S., Tadros, N. (1997). Power <strong>and</strong> Seawater Distillation From Biomass <strong>and</strong><br />

Refuse Fuels. Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong> Water<br />

Reuse. Madrid, Spain. Vol. IV, pp. 309-332<br />

Torres, C. (1991). Exergoeconomía Simbólica. Metodología para el Análisis<br />

Termoeconómico de los Sistemas Energéticos. Ph. D. Thesis. Department <strong>of</strong><br />

Mechanical Engineering. University <strong>of</strong> Zaragoza.<br />

Torres, C., Valero, A., Serra, L., Royo, J. (1999). Structural Theory <strong>and</strong><br />

<strong>Thermoeconomic</strong> Diagnosis. Part I: On Malfunction <strong>and</strong> Dysfunction Analysis.<br />

Proceedings <strong>of</strong> the ECOS’99. ASME. Tokyo, Japan. pp. 368-373.<br />

Torres, M., Medina, J. A. (1999). Desalination in Spain, a Race for Lowering Power<br />

Consumption. Proceedings <strong>of</strong> the IDA World Congress on Desalination <strong>and</strong> Water<br />

Reuse. San Diego, USA.<br />

Tribus, M., Asimow, R., Richardson, N., Gustaldo, C., Elliot, K., Chambers, J.,<br />

Evans, R. B. (1960). Thermodynamic <strong>and</strong> Economic Considerations in the<br />

Preparation <strong>of</strong> Fresh Water from the Sea. Report no. 59-34. Department <strong>of</strong><br />

Engineering. University <strong>of</strong> California.<br />

452 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


References<br />

Tribus, M., Evans, R. B. (1963). The Thermo-Economics <strong>of</strong> Sea Water Conversion.<br />

Report no. 62-53. Department <strong>of</strong> Engineering. University <strong>of</strong> California.<br />

Tsatsaronis, G., Winhold, M. (1985a). Exergoeconomic Analysis <strong>and</strong> Evaluation <strong>of</strong><br />

Energy Conversion Plants, Part I: A New General Methodology. Energy Vol. 10, pp.<br />

69-80.<br />

Tsatsaronis, G., Winhold, M. (1985b). Exergoeconomic Analysis <strong>and</strong> Evaluation <strong>of</strong><br />

Energy Conversion Plants, Part II: Analysis <strong>of</strong> a Coal Fired Steam Power Plant.<br />

Energy Vol. 10, pp. 81-94, 1985.<br />

Tsatsaronis, G. (1987). A Review <strong>of</strong> Exergoeconomic Methodologies. International<br />

Symposium on Second Law Analysis <strong>of</strong> Thermal Systems. Rome (ASME Book<br />

I00236). ASME, pp. 81-87. New York, USA.<br />

Tsatsaronis, G. (1994). Invited Papers <strong>of</strong> Exergoeconomics. Energy Vol. 19, pp. 279.<br />

Tsatsaronis, G. (1998). Recent Development in Energy Economics. Proceedings <strong>of</strong><br />

ECOS’98. Nancy, France. Vol. I, pp. 37-38.<br />

United Nations Economic <strong>and</strong> Social Commission for Western Asia (ECSWA), 1994<br />

<strong>and</strong> 1995. Private communication.<br />

University <strong>of</strong> Tennessee <strong>and</strong> Oak Ridge National Laboratory (ORNL) (1999). Netlib<br />

Repository. A Collection <strong>of</strong> Mathematical S<strong>of</strong>tware. Web page: http://www.netlib.org.<br />

Valero, A., Lozano, M. A., Muñoz, M. (1986a). A General Theory <strong>of</strong> Exergy Saving.<br />

Part I: On the Exergetic Cost. ASME Book H0341A. WAM 1986. AES-Vol 2-3,<br />

pp. 1-8.<br />

Valero, A., Lozano, M. A., Muñoz, M. (1986b). A General Theory <strong>of</strong> Exergy Saving.<br />

Part II: On the <strong>Thermoeconomic</strong> Cost. ASME Book H0341A. WAM 1986. AES-<br />

Vol. 2-3, pp. 9-16.<br />

Valero, A., Lozano, M. A., Muñoz, M. (1986c). A General Theory <strong>of</strong> Exergy Saving.<br />

Part III: Energy Saving <strong>and</strong> <strong>Thermoeconomic</strong>s. ASME Book H0341A. WAM 1986.<br />

AES-Vol 2-3, pp. 17-22.<br />

Valero, A., Torres, C. (1990). On Causality in Organized Energy Systems, Part II:<br />

Symbolic exergoeconomics. International Symposium: A future for energy. Florence,<br />

Italy. Pergamon Press, pp. 393-401.<br />

Valero, A., Serra, L., Torres, C. (1992). A General Theory <strong>of</strong> <strong>Thermoeconomic</strong>s: Part<br />

I: Structural Analysis. International Symposium ECOS’92. Zaragoza, Spain. ASME<br />

Book I00331, pp. 137-145.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 453


References<br />

Valero, A., Serra, L., Lozano, M. A. (1993). Structural Theory <strong>of</strong> <strong>Thermoeconomic</strong>s.<br />

International Symposium on Thermodynamics <strong>and</strong> the Design, Analysis <strong>and</strong><br />

Improvement <strong>of</strong> Energy Systems. Richter H. J. eds. ASME Book no. H00874. New<br />

Orleans, USA. pp. 189-198.<br />

Valero, A., Lozano, M. A., Serra, L., Torres, C. (1994). Application <strong>of</strong> the Exergy<br />

Cost Theory to the CGAM problem. Energy Vol. 19, no. 13, pp 365-381.<br />

Valero, A., Lozano, M. A. (1997). An Introduction <strong>of</strong> <strong>Thermoeconomic</strong>s. Published<br />

in Developments in the design <strong>of</strong> thermal systems (Boehm Editor). Cambridge<br />

University Press. pp. 203-233.<br />

Valero, A., Correas, L., Serra, L. (1999). On-Line <strong>Thermoeconomic</strong> Diagnosis <strong>of</strong><br />

Thermal Power Plants. Thermodynamics <strong>and</strong> Optimization <strong>of</strong> Complex Energy<br />

Systems. Klumer Academic Publishers (Bejan <strong>and</strong> Mamut eds.), pp. 117-136.<br />

Valero, A., Torres, C., Lerch, F. (1999). Structural Theory <strong>and</strong> <strong>Thermoeconomic</strong><br />

Diagnosis. Part III: Intrinsic <strong>and</strong> Induced Malfunctions. Proceedings <strong>of</strong> the<br />

ECOS’99 Conference (ASME). Tokyo, Japan. pp. 35-41.<br />

Vargaftik, N. B. (1978). H<strong>and</strong>book <strong>of</strong> Physical Properties <strong>of</strong> Liquids <strong>and</strong> Gases.<br />

Hemisphere Publishing Corporation.<br />

VA Tech Wabag (1999). List <strong>of</strong> References, Thermal Desalination Plants <strong>and</strong><br />

Reverse Osmosis Desalination. Private Communication.<br />

Villalon, C. (1995). VB Automatic Help Author Version 1.25.<br />

Von Spakovsky, M. R. (1986). A Practical Generalized Analysis Approach to the<br />

Optimal <strong>Thermoeconomic</strong> Design <strong>and</strong> Improvement <strong>of</strong> Real-World Thermal<br />

Systems. Ph. D. Thesis. Georgia Institute <strong>of</strong> Technology.<br />

Von Spakovsky, M. R., Evans, R. B. (1993). Engineering Functional Analysis.<br />

Part I. ASME Journal <strong>of</strong> energy resources technology, Vol. 115, pp. 86-92.<br />

Von Spakovsky, M. R. (1994). Application <strong>of</strong> Engineering Functional Analysis to<br />

the Analysis <strong>and</strong> Optimization <strong>of</strong> the CGAM Problem. Energy Vol. 19, no. 13.<br />

Wangnick, K. (1998). 1998 IDA Worldwide Desalting Plants Inventory Report<br />

no. 15.<br />

Water <strong>and</strong> Electricity Department (WED) <strong>of</strong> the United Arab Emirates (UAE)<br />

(1997). Private Communication.<br />

Yata, J., Minamiyama, T. (1979). An Equation for Thermal Conductivity <strong>of</strong> Water<br />

<strong>and</strong> Steam. JSME Vol. 22, no. 171, pp. 1234-1242.<br />

454 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


References<br />

Yusufova, V. D., Pepinov, R. I., Nicolayev, V. A., Zokhrabbekova, G. U., Lobcova,<br />

N. V., Tuayev, T. D. (1978). Thermophysical Properties <strong>of</strong> S<strong>of</strong>tened Seawater <strong>and</strong><br />

Salt Solutions Over a Wide Temperature <strong>and</strong> Pressure Range. Desalination 25,<br />

pp. 269-280.<br />

Zaleta, A., Ranz, L., Valero, A. (1998). Towards an Unified Measure <strong>of</strong> the<br />

Renewable Resources Availability: The Exergy Method Applied to the Water <strong>of</strong> a<br />

River. Energy Conversion <strong>and</strong> Management 39, 16-18, pp. 1911-1917.<br />

Zaleta, A. (1997). Conceptos sobre el Diagnóstico y la Evaluación Termoecónomica<br />

de Turbinas de Vapor. Ph. D. Thesis. Department <strong>of</strong> Mechanical Engineering.<br />

University <strong>of</strong> Zaragoza.<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 455


List <strong>of</strong> figures<br />

FIGURE 2.1 General outlay <strong>of</strong> MSF distillation with brine recycling ....................................................<br />

FIGURE 2.2 Flow diagram <strong>of</strong> Multi-Effect Distillation (MED) with thermal vapor compression (TVC)<br />

FIGURE 2.3 MED process with vertical tube evaporators (VTE) ..........................................................<br />

FIGURE 2.4 Flow diagram <strong>of</strong> a vapor compression system with vertical tube evaporators (VTE) .......<br />

FIGURE 2.5 Diagram model <strong>of</strong> a solar still ............................................................................................<br />

FIGURE 2.6 Reverse osmosis process.....................................................................................................<br />

FIGURE 2.7 Reverse osmosis (RO) desalination with Pelton turbine ....................................................<br />

FIGURE 2.8 Electrodialysis process........................................................................................................<br />

FIGURE 3.1 Schematic diagram <strong>of</strong> a single effect MSF evaporator with recycled brine .......................<br />

FIGURE 3.2 Cross-section <strong>of</strong> a stage in a typical MSF plant .................................................................<br />

FIGURE 3.3 Temperature pr<strong>of</strong>ile <strong>of</strong> a recycle brine MSF plant .............................................................<br />

FIGURE 3.4 A general stage in a MSF plant...........................................................................................<br />

FIGURE 3.5 Heat input section ...............................................................................................................<br />

FIGURE 3.6 Mixing <strong>and</strong> splitting points in the MSF desalination plant.................................................<br />

FIGURE 3.7 Solution algorithm <strong>of</strong> a MSF desalination plant model......................................................<br />

FIGURE 3.8 Correspondence between the Top Brine Temperature <strong>and</strong> distillate output.......................<br />

FIGURE 3.9 Brine recirculation as a function <strong>of</strong> the distillate output.....................................................<br />

FIGURE 3.10 Make-up feed water as a function <strong>of</strong> the distillate output ..................................................<br />

FIGURE 3.11 Seawater to reject section as a function <strong>of</strong> the distillate output..........................................<br />

FIGURE 3.12 Top brine temperature depending on the seawater temperature <strong>and</strong> distillate<br />

production. Data collected during the year 1997................................................................<br />

FIGURE 3.13 Recycle brine flow as a function <strong>of</strong> the seawater temperature <strong>and</strong> production.<br />

Real data collected in the MSF distillers during 1997........................................................<br />

FIGURE 3.14 Make-up feed flow obtained for each range <strong>of</strong> seawater temperature when real<br />

data are computed. Average data <strong>of</strong> 1997 ..........................................................................<br />

FIGURE 3.15 Seawater to reject flow correlations for different seawater temperatures entering<br />

the MSF plant. Data collected during the year 1997 ..........................................................<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

37<br />

39<br />

40<br />

42<br />

44<br />

46<br />

47<br />

50<br />

54<br />

55<br />

56<br />

58<br />

62<br />

63<br />

68<br />

72<br />

73<br />

73<br />

73<br />

74<br />

74<br />

75<br />

75


458<br />

List <strong>of</strong> figures<br />

FIGURE 4.1 Schematic diagram <strong>of</strong> the power generation plant. Main significant flows are<br />

numbered for later descriptions <strong>and</strong> equations ...................................................................<br />

FIGURE 4.2 Schematic diagram <strong>of</strong> a turbine section ..............................................................................<br />

FIGURE 4.3 Isoentropic <strong>and</strong> real expansion <strong>of</strong> the steam in a turbine section........................................<br />

FIGURE 4.4 TTD differences in an HP heater ........................................................................................<br />

FIGURE 4.5 TTD differences in an LP heater.........................................................................................<br />

FIGURE 4.6 Isoentropic <strong>and</strong> real compression process in a pump..........................................................<br />

FIGURE 4.7 Gl<strong>and</strong> <strong>and</strong> seal steam system ..............................................................................................<br />

FIGURE 4.8 Leakage flows <strong>and</strong> seals <strong>of</strong> a steam turbine........................................................................<br />

FIGURE 4.9 Algorithm to solve the power plant model using the Powell hybrid method .....................<br />

FIGURE 4.10 Last stage <strong>of</strong> LP turbine acting as a compressor.................................................................<br />

FIGURE 4.11 Power plant scheme in the NTW Model. Some flowstreams are renumbered<br />

with respect fig. 4.1.............................................................................................................<br />

FIGURE 5.1 SIMTAW MSF process window ........................................................................................<br />

FIGURE 5.2 SIMTAW power plant window ..........................................................................................<br />

FIGURE 6.1 Physical structure <strong>of</strong> the co-generation plant......................................................................<br />

FIGURE 6.2 Productive structure <strong>of</strong> the cogeneration plant ...................................................................<br />

FIGURE 6.3 Generic component scheme ................................................................................................<br />

FIGURE 6.4 Economic resources scheme ...............................................................................................<br />

FIGURE 6.5 Fuel / Product diagram <strong>and</strong> fuel <strong>and</strong> product exergy flows (kW) in design<br />

conditions for the co-generation plant shown in figure 6.1 ................................................<br />

FIGURE 6.6 Fuel impact <strong>and</strong> technical saving ........................................................................................<br />

FIGURE 6.7 Malfunction <strong>and</strong> fuel impact...............................................................................................<br />

FIGURE 6.8 Analysis <strong>of</strong> the irreversibility causes (kW).........................................................................<br />

FIGURE 6.9 Analysis <strong>of</strong> fuel impact (kW)..............................................................................................<br />

FIGURE 7.1 Productive structure <strong>of</strong> the simple co-generation system ...................................................<br />

FIGURE 7.2 Physical structure <strong>of</strong> the power plant considered for the thermoeconomic model .............<br />

FIGURE 7.3 Physical structure <strong>of</strong> the MSF plant considered for the thermoeconomic <strong>analysis</strong> ............<br />

FIGURE 7.4 F-P description in steam power plant..................................................................................<br />

FIGURE 7.5 Productive structure <strong>of</strong> the power plant in extraction mode ...............................................<br />

FIGURE 7.6 Changes applied to extraction mode productive structure (figure 7.5) when<br />

the plant operates in condensing mode ...............................................................................<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

78<br />

80<br />

81<br />

82<br />

84<br />

88<br />

88<br />

89<br />

92<br />

93<br />

94<br />

101<br />

103<br />

126<br />

130<br />

140<br />

142<br />

145<br />

148<br />

166<br />

150<br />

152<br />

161<br />

163<br />

165<br />

167<br />

169<br />

170


List <strong>of</strong> figures<br />

FIGURE 7.7 Productive structure corresponding to extraction mode with low energy production<br />

in a dual-purpose plant. Changes with respect to figure 7.5...............................................<br />

FIGURE 7.8 Productive structure <strong>of</strong> the steam power plant in parallel <strong>and</strong> twin extraction mode.<br />

Changes with respect to figure 7.5 .....................................................................................<br />

FIGURE 7.9 Productive structure <strong>of</strong> the steam power plant in desalination or twin desalination mode<br />

FIGURE 7.10 F-P definition in the MSF unit............................................................................................<br />

FIGURE 7.11 Productive structure <strong>of</strong> the MSF unit..................................................................................<br />

FIGURE 7.12 Physical model considered in the thermoeconomic <strong>analysis</strong> <strong>of</strong> the MSF plant .................<br />

FIGURE 7.13 Impact on fuel <strong>analysis</strong> when the efficiency <strong>of</strong> the HPT4 is decreased 10% .....................<br />

FIGURE 7.14 Irreversibility increase <strong>analysis</strong> with the inefficiency in the HPT4....................................<br />

FIGURE 7.15 Additional fuel consumption when varying the isoentropic efficiency in HPT4 ...............<br />

FIGURE 7.16 Unit electricity cost when the isoentropic HPT4 efficiency is modified............................<br />

FIGURE 7.17 Unit distilled water cost when the isoentropic HPT4 efficiency is modified .....................<br />

FIGURE 7.18 Impact on fuel <strong>analysis</strong> when the fouling in BH is neglected ........................................<br />

FIGURE 7.19 Irreversibility increase in the MSF with BH = 0. NTOS case ........................................<br />

FIGURE 7.20 Impact on fuel <strong>analysis</strong> when the fouling in heater is varied .............................................<br />

FIGURE 7.21 Monetary cost <strong>of</strong> distillate when the fouling in heater is varied.........................................<br />

FIGURE 7.22 Impact on fuel <strong>analysis</strong> without fouling in RCS. MCR case ..........................................<br />

FIGURE 7.23 Irreversibility increase <strong>analysis</strong> <strong>of</strong> section 7.3.2.3 ..........................................................<br />

FIGURE 7.24 Impact on fuel depending on fouling in recovery section ..................................................<br />

FIGURE 7.25 Monetary cost <strong>of</strong> electricity depending on the fouling in recovery section .......................<br />

FIGURE 7.26 Cost in $ per cubic meter <strong>of</strong> water when recovery section fouling is varied......................<br />

FIGURE 7.27 Impact on fuel <strong>analysis</strong> in section 7.3.2.4 .......................................................................<br />

FIGURE 7.28 Irreversibility increase in section 7.3.2.4............................................................................<br />

FIGURE 7.29 Additional fuel consumption due to inefficiencies in several components<br />

<strong>of</strong> the power plant ...............................................................................................................<br />

FIGURE 7.30 Electricity cost with five inefficiencies in the power plant ................................................<br />

FIGURE 7.31 Water cost under different degrees <strong>of</strong> inefficiency in five components .........................<br />

FIGURE 7.32 Impact on fuel <strong>analysis</strong> without fouling in distillers .........................................................<br />

FIGURE 7.33 Increase <strong>of</strong> irreversibility when fouling is neglected in MSF plant ................................<br />

FIGURE 7.34 Impact on fuel due to several inefficiencies in the MSF plant.<br />

Unit exergy cost <strong>of</strong> steam <strong>and</strong> electricity is 2.55 <strong>and</strong> 2.85 respectively.............................<br />

FIGURE 7.35 Water cost when the fouling in three distillers is varied ....................................................<br />

FIGURE 7.36 Malfunctions with an inefficiency <strong>of</strong> 5 ºC in the TTD <strong>of</strong> HPH1 heater under<br />

varying loads in the steam power plant ..............................................................................<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 459<br />

170<br />

171<br />

171<br />

172<br />

174<br />

178<br />

208<br />

208<br />

219<br />

220<br />

220<br />

232<br />

232<br />

235<br />

235<br />

246<br />

246<br />

249<br />

249<br />

250<br />

259<br />

259<br />

264<br />

264<br />

265<br />

274<br />

274<br />

277<br />

279<br />

280


460<br />

List <strong>of</strong> figures<br />

FIGURE 7.37 Malfunctions generated when the FP is working with an isoentropic efficiency<br />

12% lower than the expected under four different loads in the steam power plant ............<br />

FIGURE 7.38 Malfunctions generated by an inefficiency <strong>of</strong> 5% in the isoentropic efficiency<br />

<strong>of</strong> the HPT1 under varying loads in the steam power plant................................................<br />

FIGURE 7.39 Malfunctions generated in the fourth section <strong>of</strong> the HPT under a 10% decrease<br />

in its isoentropic efficiency.................................................................................................<br />

FIGURE 7.40 Malfunctions in LPT1 under varying loads in the steam power plant <strong>and</strong> a 15%<br />

decrease in isoentropic efficiency .......................................................................................<br />

FIGURE 7.41 Malfunctions provoked by the fouling reduction in heater at different loads.....................<br />

FIGURE 7.42 Malfunctions generated in the MSF plant at different loads with no fouling<br />

in the recovery section ........................................................................................................<br />

FIGURE 7.43 Malfunctions generated in the MSF plant when the fouling in reject section<br />

is neglected for the two analyzed loads ..............................................................................<br />

FIGURE 7.44 Impact on fuel in the MSF plant when the fouling is neglected in the three distillers.<br />

Three loads at 32 ºC seawater are included ........................................................................<br />

FIGURE 7.45 Physical model applied to the thermoeconomic optimization ............................................<br />

FIGURE 7.46 Productive structure <strong>of</strong> the thermoeconomic model applied to the<br />

thermoeconomic optimization ............................................................................................<br />

FIGURE 7.47 Optimization algorithm to find the minimum cost <strong>of</strong> the plant using local optimization...<br />

FIGURE 7.48 Speed <strong>of</strong> convergence <strong>of</strong> the local variables that are efficiencies.......................................<br />

FIGURE 7.49 Evolution <strong>of</strong> the local variables that are TTD in heaters ....................................................<br />

FIGURE 7.50 Minimization <strong>of</strong> the global cost <strong>of</strong> the system....................................................................<br />

FIGURE 7.51 Sensitivity <strong>analysis</strong> <strong>of</strong> the energetic efficiency <strong>of</strong> the boiler around<br />

the optimum point ( η = 0.8608) ........................................................................................<br />

1<br />

FIGURE 7.52 Sensitivity <strong>analysis</strong> <strong>of</strong> the efficiency <strong>of</strong> the first section <strong>of</strong> the high-pressure<br />

turbine around the optimum point ( η = 0.924)..................................................................<br />

FIGURE 7.53 Exergy cost <strong>of</strong> water (k* <strong>of</strong> steam <strong>and</strong> electricity entering the MSF is the unity),<br />

<strong>and</strong> distillate temperature at different loads at 32 ºC seawater ...........................................<br />

2<br />

FIGURE A1.1 Impact on fuel <strong>analysis</strong> with an inefficiency in HPH1 ................................................... 341<br />

FIGURE A1.2 Irreversibility <strong>analysis</strong> when the TTD in HPH1 is increased 5 ºC .................................. 341<br />

FIGURE A1.3 Impact on fuel associated with a variation in the TTD <strong>of</strong> HPH1.<br />

122 MW power plant production ........................................................................................ 343<br />

FIGURE A1.4 Cost <strong>of</strong> electricity when varying TTD in HPH1 (MCR performance case)........................ 343<br />

FIGURE A1.5 Cost <strong>of</strong> water when varying TTD in the first HPH (MCR performance case) ................... 344<br />

FIGURE A1.6 Impact on fuel <strong>analysis</strong> when a inefficiency in F in detected ......................................... 354<br />

FIGURE A1.7 Irreversibility <strong>analysis</strong> with the irreversibility in FP ...................................................... 354<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

281<br />

281<br />

282<br />

282<br />

283<br />

283<br />

284<br />

284<br />

297<br />

297<br />

304<br />

305<br />

305<br />

306<br />

309<br />

309<br />

315


List <strong>of</strong> figures<br />

FIGURE A1.8 Effect <strong>of</strong> feed pump efficiency on fuel consumption. Variational study in the<br />

MCR performance case ...................................................................................................... 355<br />

FIGURE A1.9 Effect <strong>of</strong> pump inefficiency on electricity cost (MCR performance case) ......................... 356<br />

FIGURE A1.10 Water cost when the efficiency <strong>of</strong> the feed pump is varied................................................ 356<br />

FIGURE A1.11 Impact on fuel <strong>analysis</strong> when the HPT1 efficiency is 5% less than the expected .......... 366<br />

FIGURE A1.12 Irreversibility <strong>analysis</strong> with the inefficiency in HPT1 .................................................... 366<br />

FIGURE A1.13 Model linearity with respect to an inefficiency in HPT1 ................................................... 368<br />

FIGURE A1.14 Cost <strong>of</strong> electricity depending on the degree <strong>of</strong> inefficiency applied to HPT1 (MCR case) 368<br />

FIGURE A1.15 Cost <strong>of</strong> water when the isoentropic efficiency is varied from –5% to 5% with<br />

respect to design efficiency (MCR case) ............................................................................ 369<br />

FIGURE A1.16 Impact on fuel <strong>analysis</strong>, section A1.4 ............................................................................. 379<br />

FIGURE A1.17 Irreversibility <strong>analysis</strong> in section A1.4 ........................................................................... 379<br />

FIGURE A1.18 Effect on the fuel consumption when the degree <strong>of</strong> inefficiency in the LPT<br />

is varied from the design point (MCR case)....................................................................... 380<br />

FIGURE A1.19 Cost <strong>of</strong> electricity for inefficiencies in LPT1 (MCR case)................................................. 381<br />

FIGURE A1.20 Water cost per cubic meter for inefficiencies in LPT1. 122 MW in extraction<br />

mode (MCR case) ............................................................................................................... 381<br />

FIGURE A1.21 Impact on fuel <strong>analysis</strong> in section A1.5 .......................................................................... 387<br />

FIGURE A1.22 Irreversibility increase in section A1.5 ........................................................................... 387<br />

FIGURE A1.23 Effect on fuel consumption when the fouling in recovery section is<br />

gradually decreased. 1,900 T/h <strong>and</strong> 32º C seawater ........................................................... 393<br />

FIGURE A1.24 Cost <strong>of</strong> a cubic meter <strong>of</strong> water depending on the fouling in the recovery section ............. 393<br />

FIGURE A1.25 Impact on fuel <strong>analysis</strong>, section A1.6 ............................................................................. 404<br />

FIGURE A1.26 Increase <strong>of</strong> irreversibility in section A1.6 ....................................................................... 404<br />

FIGURE A1.27 Effect on fuel consumption when the fouling in reject is varied. Nominal-temperature<br />

operation in summer (NTOS, i.e., 1,900 T/h <strong>and</strong> 32 ºC seawater temperature)................. 406<br />

FIGURE A1.28 Variation <strong>of</strong> the water cost when fouling in the reject section is decreased from<br />

the design value to zero ...................................................................................................... 407<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 461


List <strong>of</strong> tables<br />

TABLE 1.1 Ground water disposal <strong>and</strong> renewable water resources in the Gulf Countries in 1994<br />

(Alawadhi, 1999) ..............................................................................................................<br />

TABLE 1.2 Water dem<strong>and</strong> for the Gulf Countries in 1990 (ESCWA, 1994)......................................<br />

TABLE 1.3 Total installed capacity <strong>and</strong> production in the seawater desalination plant <strong>of</strong> the<br />

Gulf Area in year 1994 (Alawadi, 1999; Al-Gobaisi, 1999) ............................................<br />

TABLE 1.4 Contracted capacity <strong>of</strong> freshwater production from seawater <strong>and</strong> all waters with the<br />

existing process. The total capacity is 12.8 million cubic meters per day <strong>and</strong> 21 million<br />

cubic meters per day, respectively. Data collected in 1996 (Alawadhi, 1999) ................<br />

TABLE 1.5 Natural resources in the pacific region in the year 1998 (Goto et al., 1999)....................<br />

TABLE 1.6 Water use trends in the Pacific region (Goto et al., 1999)................................................<br />

TABLE 1.7 Desalination installations in the Pacific region. Data from 1998 (Goto et al., 1999).......<br />

TABLE 1.8 Water disposal in the African region in 1995...................................................................<br />

TABLE 1.9 Water withdrawal in North African countries. Data collected in 1990 for Algeria<br />

<strong>and</strong> Tunisia; for Egypt <strong>and</strong> Morocco data from 1992 (Al-Gobaisi, 1997) .......................<br />

TABLE 1.10 Water use in the U.S. in 1995 (Gleick, 1998)...................................................................<br />

TABLE 1.11 Desalinated water in Spain during the year 1998 (Torres <strong>and</strong> Medina, 1999) .............<br />

TABLE 1.12 Some <strong>of</strong> the RO desalination plants installed in Spain (Cadagua, 1999; Sánchez<br />

et al., 1997; Fayas <strong>and</strong> Novoa, 1997; Torres et al., 1999; AECYR, 1999) ......................<br />

TABLE 1.13 Specific consumption <strong>of</strong> the thermal desalination processes. Data obtained from<br />

several sources (Fisia-Italimpianti, 1999; I.D.E., 1999)...................................................<br />

TABLE 3.1 Fouling factors <strong>of</strong> the heat reject section in MSF Plants ..................................................<br />

TABLE 4.1 Typical x, y, <strong>and</strong> z coefficient values for the inlet TTD’s in an HP heater ......................<br />

TABLE 4.2 Typical x, y, z, a, <strong>and</strong> b coefficient values for the outlet TTD’s in an HP heater ............<br />

TABLE 4.3 Typical x, y, <strong>and</strong> z coefficient values for the inlet TTD’s in an LP heater ......................<br />

TABLE 4.4 Typical x, y, z, a, <strong>and</strong> b coefficient values for the outlet TTD’s in a LP heater...............<br />

TABLE 4.5 x, y, z, a, b, <strong>and</strong> c coefficient values in deaerator.............................................................<br />

TABLE 4.6 Values <strong>of</strong> the a coefficient for each pipe <strong>of</strong> the power model ..........................................<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

21<br />

21<br />

22<br />

23<br />

23<br />

24<br />

24<br />

25<br />

25<br />

26<br />

27<br />

28<br />

29<br />

76<br />

83<br />

83<br />

84<br />

84<br />

85<br />

87


464<br />

List <strong>of</strong> tables<br />

TABLE 4.7 Kd <strong>and</strong> Kd’ constants <strong>of</strong> the gl<strong>and</strong> <strong>and</strong> seal steam system ...............................................<br />

TABLE 4.8 Operating mode <strong>and</strong> mathematical model corresponding to the performance<br />

data cases ..........................................................................................................................<br />

TABLE 5.1 Input variables for the MCR (maximum continous rating, producing both electricity<br />

<strong>and</strong> water) case..................................................................................................................<br />

TABLE 5.2 Model validation for the MCR case..................................................................................<br />

TABLE 5.3 Input variables for the MR (maximum rating, producing only electricity)<br />

performance case ..............................................................................................................<br />

TABLE 5.4 Model validation for the MR case ....................................................................................<br />

TABLE 5.5 Input variables for the PL115 performance case (partial load with 115 MW<br />

<strong>of</strong> electricity <strong>and</strong> a heat extraction to MSF <strong>of</strong> 145 Gcal/h) ..............................................<br />

TABLE 5.6 Model validation for the PL115 performance data case ...................................................<br />

TABLE 5.7 Input variables for the PL85 performance case (partial load with 85 MW<br />

<strong>of</strong> electricity <strong>and</strong> 145 Gcal/h <strong>of</strong> extraction heat flow) ......................................................<br />

TABLE 5.8 Model validation for the PL85 performance case.............................................................<br />

TABLE 5.9 MSL2 performance case (minimum stable load with 45 MW <strong>of</strong> electricity<br />

<strong>and</strong> a <strong>combined</strong> heat extraction flow <strong>of</strong> 145 Gcal/h). Main input data ............................<br />

TABLE 5.10 Model validation for the MSL2 performance case ...........................................................<br />

TABLE 5.11 Input data <strong>of</strong> the MSL3 performance case (minimum stable load with two<br />

extractions <strong>of</strong> 150 <strong>and</strong> 145 Gcal/h to MSF units).............................................................<br />

TABLE 5.12 Model validation for the MSL3 performance case ...........................................................<br />

TABLE 5.13 Input data <strong>of</strong> the MSL4 performance case (minimum stable load with the maximum<br />

heat flow extraction to MSF unit: 170 Gcal/h) .................................................................<br />

TABLE 5.14 MSL4 performance case. Model validation......................................................................<br />

TABLE 5.15 Main input data <strong>of</strong> the ODOB case (one desalination-one boiler) ...................................<br />

TABLE 5.16 Model validation <strong>of</strong> the ODOB case.................................................................................<br />

TABLE 5.17 Main input data <strong>of</strong> the TDOB case (two desalination-one boiler)....................................<br />

TABLE 5.18 Model validation data for the TDOB case ........................................................................<br />

TABLE 5.19 Main input data <strong>of</strong> the VWO performance case (maximum capacity <strong>of</strong> the steam<br />

turbine with <strong>and</strong> extraction heat flow <strong>of</strong> 170 Gcal/h to MSF)..........................................<br />

TABLE 5.20 Model validation data for the VWO case .........................................................................<br />

TABLE 5.21 Input data <strong>of</strong> the COC performance case (boiler peak load at least 5% more than<br />

the MCR case) ..................................................................................................................<br />

TABLE 5.22 Model validation data for the COC case...........................................................................<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

89<br />

96<br />

106<br />

106<br />

107<br />

107<br />

108<br />

108<br />

109<br />

109<br />

110<br />

110<br />

111<br />

111<br />

112<br />

112<br />

113<br />

113<br />

114<br />

114<br />

115<br />

115<br />

116<br />

116


List <strong>of</strong> tables<br />

TABLE 5.23 Input data <strong>and</strong> performance parameters <strong>of</strong> the NTOS case (normal-temperature<br />

operation in summer)........................................................................................................<br />

TABLE 5.24 Model validation <strong>of</strong> the NTOS performance case ............................................................<br />

TABLE 5.25 Input data <strong>and</strong> performance parameters <strong>of</strong> the HTOS case (high-temperature<br />

operation in summer)........................................................................................................<br />

TABLE 5.26 Model validation <strong>of</strong> the HTOS performance case ............................................................<br />

TABLE 5.27 Some input data <strong>and</strong> performance parameters <strong>of</strong> the LTOS case (low-temperature<br />

operation in summer)........................................................................................................<br />

TABLE 5.28 Model validation. LTOS performance case in MSF distillers..........................................<br />

TABLE 5.29 Some input data <strong>and</strong> performance parameters <strong>of</strong> the HTOW case (high-temperature<br />

operation in winter) ..........................................................................................................<br />

TABLE 5.30 Model validation <strong>of</strong> HTOW case <strong>of</strong> the MSF plant .........................................................<br />

TABLE 6.1 Fuel <strong>and</strong> product definitions for typical dual-purpose power <strong>and</strong> desalination<br />

plant units .....................................................................................................................<br />

TABLE 6.2 Fuels <strong>and</strong> Products <strong>of</strong> the components <strong>of</strong> the co-generation plant ...................................<br />

TABLE 6.3 Characteristic equations <strong>of</strong> the cogeneration plant...........................................................<br />

TABLE 6.4 Design <strong>and</strong> operation exergy flow values <strong>of</strong> the cogeneration plant (figure 6.1) ............<br />

TABLE 6.5 Fuel/Product definition corresponding to figure 6.5 ........................................................<br />

TABLE 6.6 Increase <strong>of</strong> unit consumption. (100<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 465<br />

119<br />

119<br />

120<br />

120<br />

121<br />

121<br />

122<br />

122<br />

129<br />

131<br />

132<br />

144<br />

146<br />

∆κij)<br />

......................................................................... 146<br />

TABLE 6.7 Irreversibility matrix <strong>and</strong> unit cost <strong>of</strong> product..................................................................<br />

TABLE 6.8 Malfunction <strong>and</strong> dysfunction table in (kW) .....................................................................<br />

TABLE 7.1 Fuel, product, characteristic equation <strong>and</strong> exergy cost balance in the simple<br />

co-generation system ........................................................................................................<br />

TABLE 7.2 Results <strong>of</strong> the simple co-generation system model, MCR case........................................<br />

TABLE 7.3 Description <strong>of</strong> components appearing in figure 7.2 .........................................................<br />

TABLE 7.4 Components description from figure 7.3. Note that component no. 1 is not described<br />

in physical model but is included in other schemes .........................................................<br />

TABLE 7.5 Exergy flows <strong>and</strong> characteristic equations <strong>of</strong> components in the steam power plant<br />

(extraction mode)..............................................................................................................<br />

TABLE 7.6 Exergy flows <strong>and</strong> characteristic equations for the components <strong>of</strong> the MSF plant ...........<br />

TABLE 7.7 System <strong>of</strong> equations providing the exergy cost <strong>of</strong> the steam power plant<br />

(extraction mode)..............................................................................................................<br />

TABLE 7.8 System <strong>of</strong> equations providing the exergy costs <strong>of</strong> the MSF plant (figure 7.11) .............<br />

151<br />

152<br />

162<br />

162<br />

164<br />

165<br />

176<br />

179<br />

182<br />

184


466<br />

List <strong>of</strong> tables<br />

TABLE 7.9 Case studies <strong>of</strong> the exergy cost <strong>analysis</strong> (PTC: Performance Test Case <strong>of</strong> the dual<br />

plant; Gc: Natural gas consumed; CBS: Cleaning Ball System was used) ......................<br />

TABLE 7.10 Exergy (kW fuel/kW product) unit costs k* <strong>of</strong> most significant flows <strong>of</strong> the dual plant .<br />

TABLE 7.11 Exergoeconomic (monetary) unit costs ($/GJ) <strong>of</strong> most significant flows <strong>of</strong> a dual<br />

power <strong>and</strong> desalination plant. Cost <strong>of</strong> water c*<br />

D is expressed in $/m3,<br />

<strong>and</strong> electricity<br />

cost <strong>of</strong> is also expressed in $/kW·h (c*<br />

GEN* ) ...................................................................<br />

TABLE 7.12 Irreversibilities (exergy destruction, kW) in the different components <strong>of</strong> the dual<br />

plant. MSF unit is considered a component......................................................................<br />

TABLE 7.13 Isoentropic efficiencies <strong>of</strong> pumps <strong>and</strong> turbine sections <strong>of</strong> the power plant......................<br />

TABLE 7.14 Product <strong>and</strong> fuel (kW), <strong>and</strong> exergy efficiency (%) values for the power <strong>and</strong><br />

MS plants. Note: The efficiency <strong>of</strong> the boiler is not included in the final efficiency.......<br />

TABLE 7.15 Unit exergy costs k* (kW/kW) <strong>of</strong> component products in the steam power plant<br />

coupled with a MSF unit...................................................................................................<br />

TABLE 7.16 Costing equation parameters for an MSF <strong>and</strong> power plant (El-Sayed, 1996).<br />

Units: ca k$/ft2<br />

2<br />

, A ft , M lb/s, Q kW, Pi,<br />

Pe<br />

psia, Ti<br />

R, ∆T<br />

F, ∆P,<br />

dP psi, e = η/1–<br />

η.<br />

Subscripts: i, inlet; e, exit; t, tube; s, shell; m, mean (LTMD) ........................................<br />

TABLE 7.17 Component parameters in Boehm (1987) equations.........................................................<br />

TABLE 7.18 Costing equations proposed by Frangopoulos (1991) ......................................................<br />

TABLE 7.19 Cost equations proposed by Lozano et al. (1996). η exergetic efficiency, B exergy<br />

flow <strong>of</strong> product, S negentropy, vw velocity <strong>of</strong> tubes , W power, e eficiency <strong>of</strong> the<br />

condenser (= T0<br />

(s2<br />

– s1)/(h2<br />

– h1))<br />

...................................................................................<br />

TABLE 7.20 Price breakdown per section in a dual-purpose plant .......................................................<br />

TABLE 7.21 <strong>Thermoeconomic</strong> costs <strong>of</strong> distilled water <strong>and</strong> electricity <strong>of</strong> the analyzed<br />

dual-purpose plant.............................................................................................................<br />

TABLE 7.22 <strong>Thermoeconomic</strong> cost <strong>of</strong> electricity ($/kW·h) <strong>and</strong> water ($/m3)<br />

for the cases<br />

studied in the exergetic cost <strong>analysis</strong>................................................................................<br />

TABLE 7.23 F-P diagram in design, output power <strong>of</strong> 122 MW ........................................................... 209<br />

TABLE 7.24 F-P values with inefficiency in HPT4 (10% lower efficiency) .................................... 210<br />

TABLE 7.25 KP matrix in design (122 MW) ....................................................................................... 211<br />

TABLE 7.26 KP matrix with inefficiency in HPT4 (10%) ................................................................... 212<br />

TABLE 7.27 Variation de KP with inefficiency in HPT4...................................................................... 213<br />

TABLE 7.28 Irreversibility matrix I with an inefficiency in HPT4 ...................................................... 214<br />

TABLE 7.29 Dysfunction/malfunction matrix with inefficiency in HPT4 (10% isoentropic eff.) ....... 215<br />

TABLE 7.30 Malfunction matrix with inefficiency in HPT4 (1% isoentropic eff. is varied) ........... 216<br />

TABLE 7.31 F-P values (design) for the MSF plant. Nominal production in summer. .................... 222<br />

TABLE 7.32 F-P values without fouling in heater. Nominal production, 32 ºC seawater ................ 223<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant<br />

185<br />

186<br />

187<br />

188<br />

189<br />

190<br />

191<br />

193<br />

194<br />

194<br />

195<br />

196<br />

197<br />

197


List <strong>of</strong> tables<br />

TABLE 7.33 KP matrix in design ...................................................................................................... 224<br />

TABLE 7.34 KP matrix without fouling in heater. NTOS data case .................................................... 225<br />

TABLE 7.35 Variation <strong>of</strong> the KP matrix without fouling in heater. NTOS case .............................. 226<br />

TABLE 7.36 Irreversibility matrix without fouling in heater. 1,900 T/h <strong>and</strong> 32 ºC seawater temp ..... 227<br />

TABLE 7.37 Malfunction/dysfunction matrix without fouling in heater. NTOS case ..................... 228<br />

TABLE 7.38 Malfunction matrix varying fouling in heater 0.00001 m 2 K/W in NTOS case .......... 229<br />

TABLE 7.39 F-P values in design, 122 MW output power .................................................................. 238<br />

TABLE 7.40 F-P values without fouling in recovery section. MCR case ............................................ 239<br />

TABLE 7.41 KP matrix in design. MCR case ...................................................................................... 240<br />

TABLE 7.42 KP matrix without fouling in recovery section. MCR case ............................................ 241<br />

TABLE 7.43 Variation <strong>of</strong> KP without fouling in recovery section. MCR case .................................... 242<br />

TABLE 7.44 Irreversibility matrix without fouling in recovery section (MCR case) .......................... 243<br />

TABLE 7.45 Malfunction/dysfunction matrix without fouling in recovery section (MCR case) ......... 244<br />

TABLE 7.46 Malfunction matrix when the fouling in recovery is varied 0.00001 m 2 K/W<br />

in MCR case ................................................................................................................. 245<br />

TABLE 7.47 F-P values in design, 122 MW output power ............................................................... 252<br />

TABLE 7.48 F-P values with inefficiencies in five components (MCR case) ................................... 253<br />

TABLE 7.49 KP matrix in design (MCR Case) ............................................................................... 254<br />

TABLE 7.50 KP matrix with several inefficiencies in MCR case ..................................................... 255<br />

TABLE 7.51 Variation <strong>of</strong> KP matrix with several inefficiencies in MCR case ................................. 256<br />

TABLE 7.52 Irreversibility matrix with five inefficiencies in power plant (MCR case ............................ 257<br />

TABLE 7.53 Malfunction/dysfunction matrix with five inefficiencies in MCR case ............................... 258<br />

TABLE 7.54 Comparison <strong>of</strong> individual inefficiencies <strong>and</strong> <strong>combined</strong> inefficiencies in the<br />

power plant. The first five columns are individual inefficiencies, the sixth is<br />

the sum <strong>of</strong> the five inefficiencies <strong>and</strong> the seventh is the malfunctions generated<br />

with the five <strong>combined</strong> inefficiencies. MCR conditions .................................................. 261<br />

TABLE 7.55 Intrinsic <strong>and</strong> induced malfunctions (MF) <strong>and</strong> impact on fuel (MF*) <strong>of</strong> the power<br />

plant. 122 MW load.......................................................................................................... 262<br />

TABLE 7.56 The first column represents the X-axis in charts, corresponding to the inefficiency<br />

associated with each component ...................................................................................... 263<br />

TABLE 7.57 F-P values in design, nominal production with 32 ºC seawater ...................................... 267<br />

TABLE 7.58 F-P values without fouling in heater, recovery <strong>and</strong> reject section. NTOS case ............. 268<br />

TABLE 7.59 KP matrix in design (NTOS case) ............................................................................... 269<br />

TABLE 7.60 KP matrix with three inefficiencies in distillers (NTOS case) ...................................... 270<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 467


List <strong>of</strong> tables<br />

TABLE 7.61 Variation <strong>of</strong> KP when the fouling in distillers is zero (NTOS case). ........................... 271<br />

TABLE 7.62 Irreversibility matrix with three inefficiencies in distillers. NTOS case ...................... 272<br />

TABLE 7.63 Malfunction/dysfunction matrix when fouling in distillers is zero (NTOS case). ...... 273<br />

TABLE 7.64 Correspondence between the X-label <strong>and</strong> fouling ............................................................ 276<br />

TABLE 7.65 Comparison between the sum <strong>of</strong> individual inefficiencies <strong>and</strong> three <strong>combined</strong><br />

inefficiencies in the MSF unit. The first three columns are individual inefficiencies,<br />

the fourth is the sum <strong>of</strong> the three inefficiencies <strong>and</strong> the fifth is the malfunctions<br />

generated with the three <strong>combined</strong> inefficiencies. Nominal production with 32 ºC<br />

seawater (NTOS case) ...................................................................................................... 277<br />

TABLE 7.66 Intrinsic (MFl) <strong>and</strong> induced (MFg) malfunctions <strong>of</strong> the MSF plant <strong>and</strong> their costs<br />

(impact on fuel, MF*) under nominal production (32 ºC seawater temperature)............. 278<br />

TABLE 7.67 Impact on fuel associated with the inefficiencies in the power plant in extraction<br />

mode (MCR case) ............................................................................................................. 285<br />

TABLE 7.68 Cost variation associated with the inefficiencies in the power plant in co-generation<br />

mode (MCR) ..................................................................................................................... 286<br />

TABLE 7.69 Impact on fuel associated with the inefficiencies in the MSF plant (isolated from<br />

the power plant). 32 ºC Seawater...................................................................................... 286<br />

TABLE 7.70 Additional cost associated with the inefficiencies in the MSF plant (isolated from<br />

the power plant). 32 ºC Seawater (NTOS case)................................................................ 287<br />

TABLE 7.71 Impact on fuel associated with the inefficiencies in the MSF plant (coupled with<br />

the power plant) ................................................................................................................ 287<br />

TABLE 7.72 Additional cost associated with the inefficiencies in the MSF plant (coupled with<br />

the power plant) ................................................................................................................ 287<br />

TABLE 7.73 Intrinsic <strong>and</strong> induced malfunctions at 122 MW ............................................................... 290<br />

TABLE 7.74 Intrinsic <strong>and</strong> induced malfunctions at 140 MW ............................................................... 291<br />

TABLE 7.75 Intrinsic <strong>and</strong> induced malfunctions at 90 MW ................................................................. 292<br />

TABLE 7.76 Intrinsic <strong>and</strong> induced malfunctions at 60 MW ................................................................. 293<br />

TABLE 7.77 Intrinsic <strong>and</strong> induced malfunctions at 1,900 T/h ............................................................. 294<br />

TABLE 7.78 Intrinsic <strong>and</strong> induced malfunctions at 2,400 T/h .......................................................... 295<br />

TABLE 7.79 Resources <strong>and</strong> products in the productive structure <strong>of</strong> the thermoeconomic model.<br />

The superscript (´) is extraction mass flow rate, mdes is the steam flow to MSF unit<br />

(89.7 kg/s), D is the distilled water mass flow (2000 T/h) <strong>and</strong> b w is the exergy <strong>of</strong><br />

water leaving the MSF plant (7 kJ/kg·K).......................................................................... 298<br />

TABLE 7.80 Equations <strong>of</strong> the thermoeconomic model applied in the local optimization..................... 299<br />

TABLE 7.81 Values <strong>of</strong> parameter a in the capital cost equation <strong>of</strong> a heater ......................................... 303<br />

TABLE 7.82 Results <strong>of</strong> the local variables in the optimization process ............................................. 304<br />

TABLE 7.83 Main physical variables after the optimization process.................................................... 306<br />

468 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


List <strong>of</strong> tables<br />

TABLE 7.84 Results for the optimization <strong>of</strong> the dual-purpose plant in the MCR performance<br />

case. Exergy flows are described in figure 7.45. c is the cost (10 –6 $/kJ), with a<br />

fuel cost cf <strong>of</strong> 2×10–6 $/kJ <strong>and</strong> includes the capital cost factor kZ (kZ = ϕ·Z/P);<br />

Z (10 6 $) is the capital cost <strong>of</strong> the component.................................................................. 307<br />

TABLE 7.85 Different configurations in a dual power plant with 6 co-generation units, applied<br />

to two different power <strong>and</strong> water dem<strong>and</strong>s ...................................................................... 311<br />

TABLE 7.86 Price for water <strong>and</strong> electricity depending on the policy applied ...................................... 312<br />

TABLE 7.87 Benefit obtained in the two examples with five different price policies see<br />

previous table) .................................................................................................................. 312<br />

TABLE A1.1 F-P values in design (MCR case) ................................................................................ 333<br />

TABLE A1.2 F-P values in operation with 5 ºC TTD respect to design ............................................. 334<br />

TABLE A1.3 KP matrix in design (MCR case) ................................................................................ 335<br />

TABLE A1.4 KP matrix with inefficiency in HPH1 (MCR case) ...................................................... 336<br />

TABLE A1.5 Variation <strong>of</strong> KP matrix when TTD in the HPH1 is 5 ºC higher than the expected ........ 337<br />

TABLE A1.6 Irreversibility matrix with the inefficiency in HPH1 .................................................... 338<br />

TABLE A1.7 Malfunction/Dysfunction matrix when the TTD in HPH1 is 5 ºC higher....................... 339<br />

TABLE A1.8 Malfunction matrix when TTD in HPH1 is varied 1 ºC ................................................ 340<br />

TABLE A1.9 F-P design values ........................................................................................................ 346<br />

TABLE A1.10 F-P values with inefficiency in FP: –12% in its efficiciency ........................................... 347<br />

TABLE A1.11 KP matrix in design (MCR case) ............................................................................. 348<br />

TABLE A1.12 KP matrix when the inefficiency in FP is detected ...................................................... 349<br />

TABLE A1.13 Variation <strong>of</strong> the KP matrix when the FP is working improperly.................................... 350<br />

TABLE A1.14 Irreversibility matrix with –12% in the FP efficiency .................................................. 351<br />

TABLE A1.15 Dysfunction table <strong>and</strong> malfunction array when the FP is working with 12%<br />

lower efficiency ......................................................................................................... 352<br />

TABLE A1.16 Malfunction matrix when the efficiency <strong>of</strong> the FP varies 1% ............................................ 353<br />

TABLE A1.17 F-P values without any inefficiency. MCR case ......................................................... 358<br />

TABLE A1.18 F-P values when the HPT1 decreases 5% its efficiency (MCR case) ............................ 359<br />

TABLE A1.19 KP matrix in design (MCR case) ..................................................................................... 360<br />

TABLE A1.20 KP matrix when the inefficiency in HPT1 is 5% in its efficiency ................................. 361<br />

TABLE A1.21 Variation <strong>of</strong> the KP with the inefficiency in HPT1 (MCR case) ................................... 362<br />

TABLE A1.22 Irreversibility matrix with the inefficiency in HPT1 (MCR case) ................................. 363<br />

TABLE A1.23 Dysfunction/malfunction table when the efficiency <strong>of</strong> the HPT1 is decreased 5% ........ 364<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 469


List <strong>of</strong> tables<br />

TABLE A1.24 Malfunction matrix when the efficiency <strong>of</strong> the HPT1 is varied 1% ................................ 365<br />

TABLE A1.25 F-P values in design (MCR case) ..................................................................................... 371<br />

TABLE A1.26 F-P values with the inefficiency in LPT1, MCR case ..................................................... 372<br />

TABLE A1.27 KP matrix in design, MCR case ....................................................................................... 373<br />

TABLE A1.28 KP matrix when the efficiency in the LPT1 is decreased 15%, MCR case ..................... 374<br />

TABLE A1.29 Variation <strong>of</strong> the KP matrix with an inefficiency in LPT1, MCR case ............................. 375<br />

TABLE A1.30 Irreversibility matrix with the efficiency <strong>of</strong> the LPT1 decreased 15%, MCR case ......... 376<br />

TABLE A1.31 Dysfunction/malfunction table for an inefficiency in the LPT1 (15%), MCR case ......... 377<br />

TABLE A1.32 Malfunction matrix when the efficiency <strong>of</strong> the LPT1 is varied 1%, MCR case .............. 378<br />

TABLE A1.33 F-P values in design, NTOS case ..................................................................................... 383<br />

TABLE A1.34 F-P values with fouling in RCS=0, NTOS case .............................................................. 384<br />

TABLE A1.35 KP matrix in design, NTOS case ..................................................................................... 385<br />

TABLE A1.36 KP matrix with an inefficiency in RCS, NTOS case ....................................................... 386<br />

TABLE A1.37 Variation <strong>of</strong> the KP matrix when the fouling in RCS is neglected .................................. 389<br />

TABLE A1.38 Irreversibility matrix without fouling in RCS .................................................................. 390<br />

TABLE A1.39 Dysfunction/malfunction table without fouling in RCS, NTOS case .............................. 391<br />

TABLE A1.40 Malfunction matrix when the fouling in RCS is varied 0.00001 m 2 K/W ...................... 394<br />

TABLE A1.41 F-P values in design, NTOS case ..................................................................................... 397<br />

TABLE A1.42 F-P values when the fouling in RJS = 0, NTOS case ...................................................... 398<br />

TABLE A1.43 KP matrix in design, NTOS case ..................................................................................... 399<br />

TABLE A1.44 KP matrix with the inefficiency in RJS, NTOS case ....................................................... 400<br />

TABLE A1.45 Variation <strong>of</strong> the KP matrix when the inefficiency in RJS is detected ............................. 401<br />

TABLE A1.46 Irreversibility matrix corresponding to reject fouling in RJS, NTOS case ...................... 402<br />

TABLE A1.47 Dysfunction/malfunction table when the fouling in RJS = 0 ........................................... 403<br />

TABLE A1.48 Malfunction matrix when the fouling in RJS is varied 0.00001 m 2 K/W ..................... 408<br />

TABLE A2.1 Liquid phase composition <strong>of</strong> Reference Ambient (Szargut, 1989; Morris, <strong>and</strong><br />

Szargut, 1986)................................................................................................................... 413<br />

470 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Índex<br />

Resumen ..................................................................................................................................................... 11<br />

Abstract ..................................................................................................................................................... 15<br />

CHAPTER 1. Introduction ..................................................................................................................... 17<br />

1.1 Water requirements ................................................................................................................. 18<br />

1.2 Water quality <strong>and</strong> uses ............................................................................................................ 18<br />

1.3 World water resources <strong>and</strong> dem<strong>and</strong> ........................................................................................ 19<br />

1.3.1 Gulf Region ................................................................................................................. 19<br />

1.3.2 Pacific Region <strong>and</strong> India ............................................................................................. 23<br />

1.3.3 North Africa ................................................................................................................. 25<br />

1.3.4 US experience <strong>and</strong> the Caribbean Isl<strong>and</strong>s ................................................................... 26<br />

1.3.5 Mediterranean area <strong>and</strong> Europe ................................................................................... 27<br />

1.4 Desalination <strong>and</strong> energy .......................................................................................................... 29<br />

1.5 Why a MSF <strong>and</strong> power plant? ................................................................................................. 30<br />

1.6 <strong>Thermoeconomic</strong> <strong>analysis</strong> ....................................................................................................... 32<br />

1.7 Ph. D. Thesis development ...................................................................................................... 33<br />

CHAPTER 2. Desalination processes ..................................................................................................... 35<br />

2.1 Phase change processes: distillation <strong>and</strong> freezing ................................................................... 36<br />

2.1.1 Multi-stage flash process (MSF) ................................................................................. 36<br />

2.1.2 Multi-effect distillation (MED) ................................................................................... 38<br />

2.1.3 Vapor compression (VC) ............................................................................................. 41<br />

2.1.4 Solar distillation ........................................................................................................... 43<br />

2.1.5 Freezing process .......................................................................................................... 44<br />

2.2 Processes using membranes .................................................................................................... 45<br />

2.2.1 Reverse osmosis .......................................................................................................... 45<br />

2.2.2 Electrodialysis (ED) .................................................................................................... 49<br />

2.3 Processes acting on chemical bounds ...................................................................................... 49<br />

2.3.1 Ion exchange ................................................................................................................ 49<br />

2.4 Summary ................................................................................................................................. 51<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Índex<br />

CHAPTER 3. MSF desalination steady-state model ............................................................................ 53<br />

3.1 Process description ..................................................................................................................... 54<br />

3.2 Mathematical model <strong>of</strong> MSF unit ............................................................................................... 57<br />

3.2.1 Stage model .................................................................................................................... 58<br />

3.2.2 Brine Heater Model ........................................................................................................ 62<br />

3.2.3 Mixer <strong>and</strong> splitter model ................................................................................................. 63<br />

3.3 Auxiliary equations ..................................................................................................................... 64<br />

3.3.1 Density ............................................................................................................................ 64<br />

3.3.2 Viscosity ......................................................................................................................... 64<br />

3.3.3 Thermal conductivity ...................................................................................................... 65<br />

3.3.4 Heat capacity .................................................................................................................. 65<br />

3.3.5 Enthalpy .......................................................................................................................... 65<br />

3.3.6 Vapor pressure ................................................................................................................ 66<br />

3.3.7 Boiling point elevation ................................................................................................... 67<br />

3.3.8 Non-equilibrium allowance ............................................................................................ 67<br />

3.3.9 Demister <strong>and</strong> other losses ............................................................................................... 67<br />

3.4 Solution algorithm ...................................................................................................................... 68<br />

3.5 Simulation cases ......................................................................................................................... 70<br />

3.5.1 TBT control .................................................................................................................... 71<br />

3.5.2 Inverse problem .............................................................................................................. 71<br />

3.6 Initial data <strong>and</strong> <strong>simulation</strong> .......................................................................................................... 72<br />

3.6.1 Fouling effect .................................................................................................................. 75<br />

3.7 Summary ..................................................................................................................................... 76<br />

CHAPTER 4. Steam power plant steady-state model .......................................................................... 77<br />

4.1 Model description ....................................................................................................................... 78<br />

4.2 Mathematical model ................................................................................................................... 80<br />

4.2.1 Steam turbines ................................................................................................................ 80<br />

4.2.2 HP heat exchangers ......................................................................................................... 82<br />

4.2.3 LP heat exchangers ......................................................................................................... 83<br />

4.2.4 Deaerator ......................................................................................................................... 84<br />

4.2.5 Condenser ....................................................................................................................... 85<br />

4.2.6 Boiler .............................................................................................................................. 85<br />

4.2.7 Valves ............................................................................................................................. 86<br />

4.2.7.1 Turbine control valves ...................................................................................... 86<br />

4.2.7.2 Boiler outlet stop valve ..................................................................................... 86<br />

4.2.7.3 Boiler inlet control valve .................................................................................. 86<br />

4.2.8 Pipes ................................................................................................................................ 86<br />

4.2.9 Pumps ............................................................................................................................. 87<br />

4.2.10 Gl<strong>and</strong> <strong>and</strong> seal steam system .......................................................................................... 88<br />

4.2.11 Generator ........................................................................................................................ 89<br />

4.3 Auxiliary equations ..................................................................................................................... 90<br />

4.3.1 Thermodynamic properties ............................................................................................. 90<br />

4.3.2 Transport properties ........................................................................................................ 90<br />

472 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Índex<br />

4.4 Solution algorithm ...................................................................................................................... 90<br />

4.5 Operating modes <strong>and</strong> mathematical models ............................................................................... 92<br />

4.6 Summary ..................................................................................................................................... 96<br />

CHAPTER 5. Simulator .......................................................................................................................... 99<br />

5.1 SIMTAW structure ..................................................................................................................... 100<br />

5.2 Model validation ......................................................................................................................... 104<br />

5.2.1 Power plant ..................................................................................................................... 104<br />

5.2.1.1 MCR case .......................................................................................................... 106<br />

5.2.1.2 MR case ............................................................................................................ 107<br />

5.2.1.3 PL115 case ........................................................................................................ 108<br />

5.2.1.4 PL85 case .......................................................................................................... 109<br />

5.2.1.5 MSL2 case ......................................................................................................... 110<br />

5.2.1.6 MSL3 case ........................................................................................................ 111<br />

5.2.1.7 MSL4 case ........................................................................................................ 112<br />

5.2.1.8 ODOB case ....................................................................................................... 113<br />

5.2.1.9 TDOB case ........................................................................................................ 114<br />

5.2.1.10 VWO case ......................................................................................................... 115<br />

5.2.1.11 COC case .......................................................................................................... 116<br />

5.2.2 MSF Plant ....................................................................................................................... 117<br />

5.2.2.1 NTOS case ........................................................................................................ 119<br />

5.2.2.2 HTOS case ........................................................................................................ 120<br />

5.2.2.3 LTOS case ........................................................................................................ 121<br />

5.2.2.4 HTOW case ...................................................................................................... 122<br />

CHAPTER 6. <strong>Thermoeconomic</strong>s. Fundamentals, applications <strong>of</strong> thermoeconomic diagnosis<br />

<strong>and</strong> optimization <strong>of</strong> complex energy systems .................................................................... 123<br />

6.1 Basic concepts ............................................................................................................................ 126<br />

6.1.1 The concept <strong>of</strong> cost ......................................................................................................... 126<br />

6.1.2 Fuel, product <strong>and</strong> unit exergetic consumption ................................................................ 127<br />

6.1.3 Physical <strong>and</strong> thermoeconomic plant models ................................................................... 130<br />

6.2 Calculating thermoeconomic costs ............................................................................................. 136<br />

6.2.1 Marginal <strong>and</strong> average thermoeconomic costs ................................................................. 140<br />

6.2.2 Economic resources <strong>and</strong> thermoeconomic costs ............................................................ 142<br />

6.3 <strong>Thermoeconomic</strong> applications to thermoeconomic operation diagnosis <strong>and</strong><br />

the optimization <strong>of</strong> complex energy systems .............................................................................. 143<br />

6.3.1 Operation thermoeconomic diagnosis ............................................................................ 143<br />

6.3.1.1 Technical exergy saving ................................................................................... 144<br />

6.3.1.2 Impact on resources consumption .................................................................... 145<br />

6.3.1.3 Malfunction <strong>and</strong> dysfunction <strong>analysis</strong> .............................................................. 148<br />

6.3.1.4 Intrinsic <strong>and</strong> induced malfunctions ................................................................... 153<br />

6.3.2 <strong>Thermoeconomic</strong> optimization ....................................................................................... 155<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 473


Índex<br />

CHAPTER 7. <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant ............. 159<br />

7.1 <strong>Thermoeconomic</strong> model ............................................................................................................. 161<br />

7.1.1 A simple co-generation system ....................................................................................... 161<br />

7.1.2 Physical structure ......................................................................................................... 162<br />

7.1.3 Productive structure ........................................................................................................ 166<br />

7.1.3.1 Steam power plant ............................................................................................ 166<br />

7.1.3.2 MSF unit ........................................................................................................... 171<br />

7.1.4 <strong>Thermoeconomic</strong> model ................................................................................................. 175<br />

7.2 Cost <strong>analysis</strong> ............................................................................................................................... 180<br />

7.2.1 Exergy costs allocation ................................................................................................... 181<br />

7.2.2 Exergy cost <strong>analysis</strong> ....................................................................................................... 185<br />

7.2.3 <strong>Thermoeconomic</strong> costs ................................................................................................... 191<br />

7.2.3.1 Investment costs ............................................................................................... 192<br />

7.2.3.2 Capital costs ...................................................................................................... 197<br />

7.2.4 <strong>Thermoeconomic</strong> cost <strong>analysis</strong> ....................................................................................... 197<br />

7.2.5 Cost allocation: Indirect methods ................................................................................... 198<br />

7.2.5.1 WEA method .................................................................................................... 198<br />

7.2.5.2 Fuel cost <strong>of</strong> water in dual plants ....................................................................... 200<br />

7.3 <strong>Thermoeconomic</strong> diagnosis ........................................................................................................ 202<br />

7.3.1 <strong>Thermoeconomic</strong> diagnosis <strong>of</strong> a power <strong>and</strong> desalination plant: case studies .............. 203<br />

7.3.2 Analysis <strong>of</strong> individual inefficiencies .............................................................................. 205<br />

7.3.2.1 Inefficiency in the fourth section <strong>of</strong> the high-pressure turbine ......................... 205<br />

7.3.2.2 Using the cleaning ball system in the brine heater ........................................... 221<br />

7.3.2.3 The effect <strong>of</strong> recovery section fouling on steam power plant behavior ............ 236<br />

7.3.3 Analysis <strong>of</strong> several inefficiencies ................................................................................... 251<br />

7.3.3.1 Analysis <strong>of</strong> several simultaneous inefficiencies in the steam power plant ....... 251<br />

7.3.3.2 Analysis <strong>of</strong> several inefficiencies in the MSF plant ......................................... 265<br />

7.3.4 <strong>Thermoeconomic</strong> diagnosis <strong>and</strong> load influence in the dual plant ................................... 279<br />

7.3.4.1 Effect <strong>of</strong> inefficiencies in the power plant for different loads .......................... 280<br />

7.3.4.2 Effect <strong>of</strong> MSF unit inefficiencies under different loads ................................... 283<br />

7.3.5 Summary <strong>of</strong> applying thermoeconomic diagnosis to power <strong>and</strong> desalination plants .. 285<br />

7.4 <strong>Thermoeconomic</strong> optimization ................................................................................................... 288<br />

7.4.1 Introduction ..................................................................................................................... 288<br />

7.4.2 <strong>Thermoeconomic</strong> isolation ............................................................................................. 288<br />

7.4.3 Physical model ................................................................................................................ 296<br />

7.4.4 <strong>Thermoeconomic</strong> model ................................................................................................. 296<br />

7.4.5 Local <strong>and</strong> global variables .............................................................................................. 299<br />

7.4.6 Local optimization <strong>of</strong> subsystems .................................................................................. 301<br />

7.4.7 Local optimization results ............................................................................................... 303<br />

7.5 Economic <strong>analysis</strong>. Cost, price <strong>and</strong> benefit ................................................................................ 309<br />

7.5.1 Case study ....................................................................................................................... 311<br />

7.6 Conclusions <strong>and</strong> operation recommendations ............................................................................ 313<br />

7.6.1 Cost <strong>analysis</strong> ................................................................................................................... 313<br />

7.6.1.1 Results .............................................................................................................. 313<br />

7.6.1.2 Conclusions <strong>and</strong> operation recommendations .................................................. 316<br />

474 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant


Índex<br />

7.6.2 <strong>Thermoeconomic</strong> diagnosis ............................................................................................ 317<br />

7.6.2.1 Results .............................................................................................................. 317<br />

7.6.2.2 Conclusions <strong>and</strong> final considerations ............................................................... 318<br />

7.6.3 Local optimization .......................................................................................................... 321<br />

7.6.4 Operating management ................................................................................................... 322<br />

CHAPTER 8. Synthesis, contributions <strong>and</strong> perspectives ..................................................................... 323<br />

8.1 Synthesis ..................................................................................................................................... 323<br />

8.2 Main contributions ..................................................................................................................... 325<br />

8.2.1 Simulator <strong>of</strong> a dual-purpose power <strong>and</strong> desalination plant ............................................ 325<br />

8.2.2 State <strong>of</strong> the art in <strong>Thermoeconomic</strong>s .............................................................................. 325<br />

8.2.3 F-P definition for a MSF unit ......................................................................................... 326<br />

8.2.4 Cost <strong>analysis</strong> <strong>of</strong> a dual-plant ........................................................................................... 326<br />

8.2.5 Diagnosis <strong>of</strong> a complex system ...................................................................................... 326<br />

8.2.6 Local optimization <strong>of</strong> the steam power plant ................................................................. 327<br />

8.2.7 Cost, price <strong>and</strong> benefit .................................................................................................... 327<br />

8.3 Perspectives ................................................................................................................................ 327<br />

8.3.1 Improving existing plants. Process integration .............................................................. 327<br />

8.3.2 Improvements in thermoeconomic diagnosis ................................................................. 328<br />

8.3.3 Integrating attitudes ........................................................................................................ 329<br />

8.3.4 Sustainable desalination ................................................................................................. 329<br />

8.3.5 Promote energy <strong>and</strong> water interactions .......................................................................... 330<br />

ANNEX 1. <strong>Thermoeconomic</strong> diagnosis .................................................................................................. 331<br />

A1.1 Effect <strong>of</strong> an inefficiency in the high-pressure heater no.1 (HPH1) ............................................ 332<br />

A1.2 Effect <strong>of</strong> feed pump isoentropic efficiency ................................................................................ 345<br />

A1.3 Effect <strong>of</strong> an inefficiency in the first section <strong>of</strong> the high-pressure turbine (HPT1) ..................... 357<br />

A1.4 Effect <strong>of</strong> inefficiency in the first section <strong>of</strong> the low-pressure turbine (LPT1) ........................... 369<br />

A1.5 Effect <strong>of</strong> the cleaning ball system in the recovery section ......................................................... 382<br />

A1.6 Effect <strong>of</strong> reject section fouling ................................................................................................... 395<br />

A1.7 Summary .................................................................................................................................... 407<br />

ANNEX 2. Thermodynamic properties <strong>of</strong> seawater ............................................................................. 409<br />

A2.1 Specific enthalpy h <strong>of</strong> superheated or saturated vapor ............................................................ 419<br />

A2.2 Specific entropy <strong>of</strong> superheated or saturated vapor ................................................................ 410<br />

A2.3 Specific volume <strong>of</strong> superheated or saturated vapor ................................................................. 411<br />

A2.4 Latent heat vaporisation <strong>of</strong> water as a function <strong>of</strong> boiling temperature .................................. 411<br />

A2.5 Seawater exergy ......................................................................................................................... 412<br />

A2.5.1 Theory ............................................................................................................................. 430<br />

A2.5.2 Practice: Brine exergy as a function <strong>of</strong> temperature, pressure <strong>and</strong> salt concentration 417<br />

<strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant 475


Índex<br />

ANNEX 3. Technical data ....................................................................................................................... 419<br />

A3.1 MSF plant ................................................................................................................................ 419<br />

A3.2 Power Plant .............................................................................................................................. 426<br />

NOMENCLATURE ........................................................................................................................................ 433<br />

REFERENCES ................................................................................................................................................ 441<br />

LIST OF FIGURES.......................................................................................................................................... 457<br />

LIST OF TABLES .......................................................................................................................................... 463<br />

476 <strong>Thermoeconomic</strong> <strong>analysis</strong> <strong>and</strong> <strong>simulation</strong> <strong>of</strong> a <strong>combined</strong> power <strong>and</strong> desalination plant

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!