Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Advanced Power Generation Systems
Advanced Power Generation Systems
Advanced Power Generation Systems
Ebook1,297 pages19 hours

Advanced Power Generation Systems

Rating: 5 out of 5 stars

5/5

()

Read preview

About this ebook

Advanced Power Generation Systems examines the full range of advanced multiple output thermodynamic cycles that can enable more sustainable and efficient power production from traditional methods, as well as driving the significant gains available from renewable sources. These advanced cycles can harness the by-products of one power generation effort, such as electricity production, to simultaneously create additional energy outputs, such as heat or refrigeration. Gas turbine-based, and industrial waste heat recovery-based combined, cogeneration, and trigeneration cycles are considered in depth, along with Syngas combustion engines, hybrid SOFC/gas turbine engines, and other thermodynamically efficient and environmentally conscious generation technologies. The uses of solar power, biomass, hydrogen, and fuel cells in advanced power generation are considered, within both hybrid and dedicated systems.

The detailed energy and exergy analysis of each type of system provided by globally recognized author Dr. Ibrahim Dincer will inform effective and efficient design choices, while emphasizing the pivotal role of new methodologies and models for performance assessment of existing systems. This unique resource gathers information from thermodynamics, fluid mechanics, heat transfer, and energy system design to provide a single-source guide to solving practical power engineering problems.

  • The only complete source of info on the whole array of multiple output thermodynamic cycles, covering all the design options for environmentally-conscious combined production of electric power, heat, and refrigeration
  • Offers crucial instruction on realizing more efficiency in traditional power generation systems, and on implementing renewable technologies, including solar, hydrogen, fuel cells, and biomass
  • Each cycle description clarified through schematic diagrams, and linked to sustainable development scenarios through detailed energy, exergy, and efficiency analyses
  • Case studies and examples demonstrate how novel systems and performance assessment methods function in practice
LanguageEnglish
Release dateJul 15, 2014
ISBN9780123838612
Advanced Power Generation Systems
Author

Ibrahim Dincer

Dr. Ibrahim Dincer is professor of Mechanical Engineering at the Ontario Tech. University and visiting professor at Yildiz Technical University. He has authored numerous books and book chapters, and many refereed journal and conference papers. He has chaired many national and international conferences, symposia, workshops, and technical meetings. He has also delivered many plenary, keynote and invited lectures. He is an active member of various international scientific organizations and societies, and serves as editor in chief, associate editor, regional editor, and editorial board member for various prestigious international journals. He is a recipient of several research, teaching and service awards, including the Premier?s Research Excellence Award in Ontario, Canada. For the past seven years in a row he has been recognized by Thomson Reuters as one of The Most Influential Scientific Minds in Engineering and one of the Most Highly Cited Researchers.

Read more from Ibrahim Dincer

Related to Advanced Power Generation Systems

Related ebooks

Power Resources For You

View More

Related articles

Reviews for Advanced Power Generation Systems

Rating: 5 out of 5 stars
5/5

2 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Advanced Power Generation Systems - Ibrahim Dincer

    Advanced Power Generation Systems

    First Edition

    Ibrahim Dincer

    Calin Zamfirescu

    University of Ontario Institute of Technology, 2000 Simcoe St N, Oshawa, Ontario L1H 7K4, Canada

    Table of Contents

    Cover image

    Title page

    Copyright

    Acknowledgments

    Preface

    Chapter 1: Fundamentals of Thermodynamics

    Abstract

    1.1 Introduction

    1.2 Thermodynamic Properties and Basic Concepts

    1.3 Equations of State and Ideal Gas Behavior

    1.4 Laws of Thermodynamics

    1.5 Exergy

    1.6 Balance Equations for Thermodynamic Analysis

    1.7 Efficiency Definitions

    1.8 Concluding Remarks

    Chapter 2: Energy, Environment, and Sustainable Development

    Abstract

    2.1 Introduction

    2.2 Energy Resources Available on Earth

    2.3 Environmental Impact of Power Generation Systems

    2.4 Sustainability Assessment of Power Generation Technologies

    2.5 Concluding Remarks

    Chapter 3: Fossil Fuels and Alternatives

    Abstract

    3.1 Introduction

    3.2 Fuels Classification and Main Properties

    3.3 Fossil Fuels

    3.4 Alternative Fuels

    3.5 Concluding Remarks

    Chapter 4: Hydrogen and Fuel Cell Systems

    Abstract

    4.1 Introduction

    4.2 Hydrogen

    4.3 Hydrogen Production Methods

    4.4 Fuel Cells

    4.5 Fuel Cell Modeling

    4.6 Optimization of Fuel Cell Systems

    4.7 Integrated Fuel Cell Systems for Power Generation

    4.8 Concluding Remarks

    Chapter 5: Conventional Power Generating Systems

    Abstract

    5.1 Introduction

    5.2 Vapor Power Cycles

    5.3 Gas Power Cycles

    5.4 Combined Cycle Power Plants

    5.5 Hydropower Plants

    5.6 Concluding Remarks

    Chapter 6: Nuclear Power Generation

    Abstract

    6.1 Introduction

    6.2 Nuclear Reactions

    6.3 Nuclear Fuel

    6.4 Nuclear Reactors

    6.5 Nuclear-Based Cogeneration Systems

    6.6 Concluding Remarks

    Chapter 7: Renewable-Energy-Based Power Generating Systems

    Abstract

    7.1 Introduction

    7.2 Solar Power Generation Systems

    7.3 Wind Energy Systems

    7.4 Geothermal Power Generation Systems

    7.5 Biomass Energy Systems

    7.6 Ocean Energy Systems

    7.7 Concluding Remarks

    Chapter 8: Integrated Power Generating Systems

    Abstract

    8.1 Introduction

    8.2 Multistaged Systems

    8.3 Cascaded Systems

    8.4 Combined Systems

    8.5 Hybrid Systems

    8.6 Case Studies

    8.7 Concluding Remarks

    Chapter 9: Multigeneration Systems

    Abstract

    9.1 Introduction

    9.2 Key Processes and Subsystems for Multigeneration

    9.3 Assessment and Optimization of Multigeneration Systems

    9.4 Case Studies

    9.5 Concluding Remarks

    Chapter 10: Novel Power Generating Systems

    Abstract

    10.1 Introduction

    10.2 Novel Ammonia–Water Power Cycles

    10.3 Solar Thermoelectrical Power Generation

    10.4 Chemical Looping Combustion for Power Generation

    10.5 Linear Engine Power Generators

    10.6 Concluding Remarks

    Appendix A: Conversion Factors

    Appendix B: Thermophysical Properties

    Index

    Copyright

    Elsevier

    Radarweg 29, PO Box 211, 1000 AE Amsterdam, Netherlands

    The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, UK

    225 Wyman Street, Waltham, MA 02451, USA

    First edition 2014

    Copyright © 2014 Elsevier Inc. All rights reserved.

    No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means electronic, mechanical, photocopying, recording or otherwise without the prior written permission of the publisher.

    Permissions may be sought directly from Elsevier's Science & Technology Rights Department in Oxford, UK: phone (+ 44) (0) 1865 843830; fax (+ 44) (0) 1865 853333; email: permissions@elsevier.com. Alternatively you can submit your request online by visiting the Elsevier web site at http://elsevier.com/locate/permissions, and selecting Obtaining permission to use Elsevier material

    Notice

    No responsibility is assumed by the publisher for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions or ideas contained in the material herein. Because of rapid advances in the medical sciences, in particular, independent verification of diagnoses and drug dosages should be made.

    Library of Congress Cataloging-in-Publication Data

    A catalog record for this book is available from the Library of Congress

    British Library Cataloguing in Publication Data

    A catalogue record for this book is available from the British Library

    For information on all Elsevier publications

    visit our web site at store.elsevier.com

    Printed and bound in USA

    14 15 16 17 18  10 9 8 7 6 5 4 3 2 1

    ISBN: 978-0-12-383860-5

    Acknowledgments

    Ibrahim Dincer; Calin Zamfirescu, Oshawa, June 2014

    Bringing a book to fruition is not an easy job since it requires lots of effort and time tirelessly put upon. This particular book on Advanced Power Generation has caused many sleepless nights, without tears, but with smiles. We have extensively benefited from our previously published works, especially with the graduate students of Dr. Ibrahim Dincer, namely, Pouria Ahmadi, Eda Cetinkaya, Rami El-Emam, Ahmet Ozbilen, Hasan Ozcan and Tahir Ratlamwala for various illustrations and case studies. We sincerely acknowledge their assistance. Dr. Dincer also acknowledges the support provided by the Turkish Academy of Sciences.

    Last but not least, we warmly thank our wives, Gulsen Dincer and Iuliana Zamfirescu, and our children Meliha, Miray, Ibrahim Eren, Zeynep and Ibrahim Emir Dincer, and Ioana Zamfirescu. They have been a great source of support and motivation, and their patience and understanding throughout this book have been most appreciated.

    Preface

    Ibrahim Dincer; Calin Zamfirescu, Oshawa, June 2014

    Energy has been a critical issue for humankind over the centuries, and it has shaped the past, is shaping the present, and will definitely shape the future. Our existence depends essentially on it and the way we generate, convert, transform, transport, and utilize it. The picture has become even more complicated with environmental consequences and sustainability issues. When we look at the dimensions of energy, power generation stands out as the most crucial commodity and is the main drive behind the economies.

    Due to increasing energy, environmental, and sustainability issues, we need to go beyond conventional practices and hence conventional power production technologies, systems, and applications. It has become a kind of ultimate target to make power generation systems more efficient, cost effective, and more environmentally friendly and to develop potential options to fulfill these requirements. This book offers a unique coverage of the conventional and novel power generating systems in a harmonized manner with renewables and includes our recent developments on system integration and multigeneration. It is expected to serve the power generation community by providing comprehensive tools for system design, analysis, assessment, and improvement.

    Being a relatively new, the emerging technology of power generation using advanced thermodynamic cycles is found as subject matter scattered in specialized research journals and a few books, written at an advanced level and devoted exclusively to particular multiple output thermodynamic cycles. There is a scarcity of publications that introduce all these cycles for power generation in a single volume to a beginner. Therefore, the proposed book will fill the void of a much needed and relatively inexpensive textbook at the post-graduate level in the area of power generation. While teaching a particular course on power generation cycles to post-graduate engineering students, the authors themselves felt this need, which inspired them to write the proposed book.

    This book is essentially intended to serve a research-oriented textbook with a comprehensive coverage of fundamentals and main concepts, which can be used for system design, analysis, assessment, and improvement, and it therefore includes practical features in a usable format often not included in other solely academic textbooks. There is also room for this book to be used by senior undergraduate and graduate students in mainstream engineering fields (such as mechanical and electrical engineering) and new engineering programs on energy and/or power. Due to its extensive coverage it can also serve as a primary reference book for practicing engineers and researchers.

    The book consists of ten chapters which nicely amalgamate all power-related aspects, ranging from basic fundamentals to novel power generating systems, by considering efficiency, environment, and sustainability issues. The introductory chapter addresses general concepts, fundamental principles, and basic aspects of thermodynamics, energy, entropy, and exergy. These topics are covered in a broad manner, so as to furnish the reader with the background information necessary for subsequent chapters. Chapter 2 provides detailed information on energy, environment, and sustainable development, while Chapter 3 focuses on traditional fossil fuels and potential alternative fuels or fuel sources. Chapter 4 gives a distinct coverage on hydrogen and fuel cells and their systems for power generation. Chapter 5 dwells on conventional power generating systems and their design, analysis, assessment, and improvement. Chapter 6 is about nuclear power for power production with a comparative analysis and assessment, while Chapter 7 takes a comprehensive focus on renewable energy-based systems and applications for power production. Chapter 8 offers advanced power generating systems with illustrative examples and case studies. Chapter 9 delves into integrated multigeneration systems for power production, and Chapter 10 makes a good close with descriptions of novel power generating systems from our own works and some literature studies. Incorporated throughout are many illustrative examples and case studies, which provide the reader with a substantial learning experience, especially in areas of practical application. Complete references are included to point the truly curious reader in the right direction. Information on topics not covered fully in the text can, therefore, be easily found.

    We hope this book brings a new dimension to power generating practices and helps the community implement better solutions for a better future.

    Chapter 1

    Fundamentals of Thermodynamics

    Abstract

    In this chapter, thermodynamic fundamentals are revisited from the perspective of energy system analysis and assessment. In the first section, the main thermodynamic concepts and properties are introduced and defined. Basic properties such as temperature, pressure, and specific volume (measurable quantities) and concepts such as thermodynamic system, work, heat, energy, and equilibrium are discussed in detail. A separate section outlines the importance of state equations and the behavior of ideal gas during basic processes such as expansion, compression, and isochoric transformation. Thermodynamic laws and the Carnot cycle are treated in a dedicated section in which the concept of irreversibility, entropy, exergy, and exergy destruction are introduced. Subsequently thermodynamic analysis and design are introduced using mass, energy, entropy, and exergy balance equations. The seventh section of the chapter is dedicated to efficiency definitions which are tools used for assessment of power generation systems. Attention is paid to efficiency definition of systems that generate multiple outputs such as power and heat. Exergy, not energy, is the standard used to express system output. The chapter includes a section on thermodynamic analysis of psychrometric systems. These are important for many power generation units such as air-breathing engines.

    Keywords

    Energy

    Exergy

    Efficiency

    Thermodynamic properties

    State equations

    Laws of thermodynamics

    Work

    Heat

    Entropy

    Exergy

    Psychometrics

    Nomenclature

    A   area, m²

    a   acceleration, m/s²

    C   specific heat, kJ/kgK

    COP   coefficient of performance

    DOF   degree of freedom

      energy rate, kW

    e   specific total energy, kJ/kg

      total specific exergy, kW

    ex   specific exergy, kJ/kg

    F   force, N

    g   Gibbs free energy, kJ/kg

    g   gravity acceleration, m/s²

    h   specific enthalpy, kJ/kg

    KE   kinetic energy, kJ

    kB   Boltzmann constant

    l   displacement, m

    M   molecular mass, kg/kmol

    m   mass, kg

    N   number of particles

    n   number of moles, mol

    NA   number of Avogadro, molecules/kmol

    P   pressure, Pa

    p   probability

    PE   potential energy, kJ

    Q   heat, kJ

      heat rate, kW

    q   mass specific heat, kJ/kg

      universal gas constant, kJ/kmolK

    R   gas constant, kJ/kgK

    S   entropy, J/K

    s   specific entropy, kJ/kgK

      entropy rate, W/K

    T   temperature, K

    t   time, s

    U   internal energy, kJ

    u   specific internal energy, kJ/kg

    V   volume, m³

      velocity, m/s

    v   specific volume, m³/kg

    W   work, J

      work rate, kW

    w   mass specific work, kJ/kg

    x   vapor quality

    Z   compressibility factor

    z   elevation, m

    Greek letters

    η   energy efficiency

    γ   adiabatic expansion coefficient

    μ   chemical potential, kJ/kg

    ω   humidity ratio

    ϕ   total specific flow energy, kJ/kg

    φ   relative humidity

    Ψ   exergy efficiency

    θ   total specific non-flow energy, kJ/kg

    Subscripts

    0   reference state

    b   boundary work

    C   cold

    ch   chemical

    conc   concentration

    deliv   delivered

    f   flow work

    gen   generated

    H   hot

    in   input

    ke   kinetic energy

    L   low

    loss   loss

    nf   non-flow

    out   output

    p   constant pressure

    pe   potential energy

    ph   physical

    r   reduced value

    rev   reversible

    surr   surroundings

    sys   system

    u   useful work

    v   constant volume

    Superscripts

      saturated liquid

      saturated vapor

    ch   chemical

    1.1 Introduction

    In the modern world electric power is a widespread and intensely used commodity. Electric power is in demand almost everywhere: it powers appliances, provides lighting, moves vehicles, runs machines, powers computers, and the internet, and it is widely used in industrial processes to produce all sorts of goods and commodities.

    Electricity appears to be the most versatile form of energy transport because it can be converted into many other forms without major loss and used in many ways at a competitive cost. For example, electric power can be converted completely into heat using electric heaters. Also, electric power can be converted into mechanical (motive) power with a very high efficiency using electric motors. Moreover, electric power can be converted into light using devices such as light emitting diodes and incandescent lamps. Electric power can also be used to convert water into hydrogen and oxygen by electrolysis producing a fuel, a substance which stores energy in chemical form. Other chemicals can be produced electrochemically using energy generated by electric power.

    The high efficiency of electric power conversion is explained by the flow of electric charges. In physics, power is defined as the rate of energy transfer. Electric power represents the rate of transfer of electrical energy due to movement of electric charges (generally electrons). This flow of electrons cannot be observed by the naked eye. Flow of electricity through a wire can be compared in general terms to the flow of water through a duct. Similarly, flow of heat is a thermal energy transfer due to a gradient of temperature.

    Thermodynamics differentiates two kinds of energy: (i) organized (such as mechanical, electrical or electromagnetic, photonic, gravitational etc.), and (ii) disorganized (thermal energy or heat). According to the second law of thermodynamics (SLT), which will be detailed further in this chapter, organized forms of energy can be completely converted into any other form of energy. However, thermal energy cannot be fully converted into organized forms of energy due to intrinsic irreversibilities.

    Thermodynamics comes from therme which means heat in Greek and dynamis, the Greek word for power. Therefore, a common example is given as the conversion of thermal energy into mechanical energy. Thermodynamics is commonly defined as the science of energy and entropy. However, we find this partly incomplete and inconsistent and modify it as thermodynamics is the science of energy and exergy. By this definition we can compare both energy and exergy since they possess the same units and can correspondingly be used for performance assessment and comparison of thermodynamic systems.

    Any power generation system must be designed in accordance to the principles of thermodynamics. Most traditional power generation systems use a combustion process to generate high-temperature heat from a conventional fuel (coal, petroleum, natural gas, or uranium). Then, using a power cycle, thermal energy is converted into the motive power of a rotating shaft. The generated mechanical energy is then converted into electrical energy by an electric generator, producing electric power.

    Novel power generation devices such as fuel cells are able to directly convert the chemical energy of a fuel into electricity. Typically, a fuel cell consumes a synthetic fuel such as hydrogen or methanol, or it may consume a fossil fuel such as natural gas. Electrochemical reduction and oxidation reactions occur in a fuel cell generating electric power. In addition, multi-output-generation systems are currently in development throughout the world. These systems can generate multiple outputs such as electric power, heating, cooling, hydrogen, drinking water, and synthetic fuels.

    In this chapter, concepts and methods of thermodynamics are introduced for use in the design, analysis and assessment of power generation systems. The first and second laws of thermodynamics are introduced and discussed, and the use of mass, energy, entropy and exergy balance equations are written and explained in detail. The efficiency of thermodynamic systems, devices, and processes is defined in relation to energy and exergy, and various relevant examples are given.

    1.2 Thermodynamic Properties and Basic Concepts

    In this section the most important thermodynamic concepts and thermodynamic properties are introduced. The fundamental concept of thermodynamics is the thermodynamic system proposed by Carnot (1824). By definition, a thermodynamic system is separated by a real or imaginary boundary from the rest of the universe, denoted as the surroundings. For the purpose of a thermodynamic analysis, the universe is divided into two parts, the thermodynamic system and the surroundings. A thermodynamic system that can exchange energy but does not exchange matter (mass) with its surroundings is said to be a closed thermodynamic system or control mass. If a closed thermodynamic system does not exchange energy in any form with its surroundings then the system is an isolated system. If a thermodynamic system can interact with its surroundings by mass transfer and energy transfer then the system is denoted as an open system or control volume. Two main forms of energy transfer exist; these are work and heat and will be discussed in detail later. Figure 1.1 illustrates the two types of systems: (a) a closed system and (b) an open system. Figure 1.2 depicts an insulated thermodynamic system.

    Figure 1.1 Thermodynamic systems: (a) a closed system, (b) an open system.

    Figure 1.2 Isolated system—a particular type of closed system.

    The universe represents the cosmic system, consisting of matter and energy. In addition, matter is defined as particles with rest mass, including quarks and leptons, which are able to combine and form electrons, protons and neutrons. Chemical elements consist of matters formed by combinations of protons, neutrons, and electrons, whereas electrons are a class of leptons.

    The interaction between a thermodynamic system and its surroundings is affected by an intermediary of forces or force fields. A field represents a region of the universe affected by forces of a given kind. Consequently, it is possible for a field to have no matter. However, the contrary is not true. A field may exist anywhere in the universe even where matter (such as particles with rest mass) is nonexistent.

    A particular thermodynamic system is the ideal vacuum characterized by the lack of any form of matter but the presence of fields. Some examples of fields are gravitational, magnetic, and electric. Bulk matter or substance is formed by groups of atoms, molecules, and clusters of atoms and molecules. There are four forms of aggregation of substances, solid, liquid, gas, and plasma.

    In physics there are seven fundamental quantities, and their units are defined according to the International System of Units (SI). All other quantities and units—denoted as derived—can be determined from these fundamental quantities based on physical laws. Not all fundamental quantities of the SI system are directly relevant to thermodynamics. For example, luminous intensity may not be very important in thermodynamics when compared to temperature. However, derived quantities such as specific volume and pressure are very important. Table 1.1 presents the physical quantities used in thermodynamics which are directly measurable. Measurable quantities are important because they can offer quantitative information regarding the thermodynamic state of a system. Other quantities can be determined based on measurable quantities and relevant equations. Table 1.2 presents a number of physical constants relevant in thermodynamics. The table shows the constant name, its symbol, its value and measurable unit, and a brief definition. Some constants are fundamental constants of physics (e.g., Planck’s constant) or standard parameters (e.g., standard atmospheric pressure).

    Table 1.1

    Main Measurable Quantities Used in Thermodynamics

    Sources: ISU (2006).

    Table 1.2

    Some Important Constants in Thermodynamics

    Some important physical properties derived from the measurable quantities listed in Table 1.1 are introduced next. The amount of matter (or the number of moles) represents the number of unambiguously specified entities of a substance such as electrons, atoms, or molecules. The unit of the amount of matter is a mol. According to ISU (2006), this represents the amount of substance in a system which contains as many elementary entities as there are atoms in 0.012 kg of ¹²C. Another important property is the molecular mass of a substance, which represents the ratio between the mass and the amount of matter, namely

       (1.1)

    where n represents the number of moles (or the amount of matter).

    The action of forces produces changes in the magnitude and/or direction of velocity of a particular piece of matter. The magnitude of force is proportional to the mass of the matter on which it acts and to acceleration, which is defined as the variation of velocity the position vector of a piece of matter (material point) then velocity is expressed by the derivative

       (1.2)

    and the acceleration is defined by

       (1.3)

    therefore, the vector force is expressed by the following equation

       (1.4)

    The total force exerted on a surface is denoted as pressure. By definition

    where F is the magnitude of force in the direction perpendicular to the surface. The unit for pressure is Pa (Pascals), where 1 Pa = 1 N/m².

    Often in thermodynamics one deals with the concept of flow, which represents continuous movement of substance along a stream. In addition to carrying matter, a flow can transport energy and do useful work. In ) is defined as the rate of mass transport at a cross section. Specific volume (v, in m³/kg) is the parameter that connects the volumetric and mass flow rates.

    Table 1.3

    Definitions Table of Some Important Notions in Thermodynamics

    )

    Consequently, work is proportional to the magnitude of force and the magnitude of displacement, but also depends on the relative direction of the force and the displacement. For example, if the displacement direction is perpendicular to the direction of force it can be inferred that this displacement is not produced by the respective force but has other causes; in this case, the work is nil. On the contrary, if force and displacement have the same direction, then the work of the force is the maximum possible and equal to the force magnitude and the displacement. Table 1.3 (item #5) illustrates the work of gravitational force which is developed when a system is displaced in the gravitational field. When the displacement is in a vertical direction then the force and displacement vectors act in the same direction, therefore the work is

       (1.5)

    where z represents the displacement with respect to a reference system.

    If the boundary of a thermodynamic system or control volume moves, then it produces or consumes work because it displaces matter. The work done by boundary movement is referred to as boundary work. As illustrated in the Table 1.3 (item #6), boundary work is expressed by

       (1.6)

    It is interesting to remark that the boundary work for the process 1–2 equals the area of the process path represented by the Pv diagram (see Table 1.3, #6). A process path is a graphical representation of all thermodynamic states through which a system passes during a transformation. In a cyclical process for which the path 2–1 is different from the path 1–2 as illustrated in the Pv diagram, the system produces or consumes work continuously. The net work exchanged with the surroundings (Wnet) is represented by the area enclosed by the cycle. Note that this area is positive if work is delivered to surroundings and negative if work is supplied to the system from the surroundings. This observation is the basis of sign convention in thermodynamics which will be explained later.

    Furthermore, in order for a system to exchange work with its surroundings it must be either a closed system with a movable boundary or an open system (control volume). Isolated systems cannot exchange work or any kind of energy with their surroundings.

    Another form of work defined in Table 1.3 is the flow work which is characteristic of control volumes only. Flow work is a measure of the action of pushing (or pulling) a volume of fluid inside a control volume. This represents a flow-displacement work. The flow work is equal to the product of pressure and volume. Specifically, it becomes

       (1.7)

    or, in differential form, dwf = d(P v).

    Many devices such as turbines, compressors, and pumps are open systems that perform or consume work. Both flow work and boundary work are exchanged when operating these devices. The work generated or consumed by an open system (control volume) is denoted as useful work and represents the difference between boundary work and flow work: Wu = Wb − Wf. In differential form, it is written as

       (1.8)

    therefore, the useful work becomes

    The ability to do work quantifies a change in energy which is a scalar quantity evidenced by indirect measurement of energy changes caused by various interactions of a thermodynamic system with its surroundings. There are several forms of energy that fall into two categories: microscopic and macroscopic. There are two basic energy exchange (transfer) mechanisms, namely work and heat. As described above, work refers to that energy transfer which takes place by displacement of matter, either microscale or macroscale. Heat transfer is a different mechanism of energy exchange related to kinetic energy of the molecules or atoms. Heat amount, Q, is measured with the same unit as work and energy, the Joule. The energy exchange by heat will be detailed later.

    is

    When a body moves in a force field, its total macroscopic energy is a sum of kinetic and potential energy. When the change in associated potential energy of displaced matter does not depend on the path taken but only on the initial and final positions of the displacement, the force field is denoted as conservative. Any conservative force field has an associated potential energy. There are many types of potential energies, each with an associated type of force field:

    • gravitational,

    • elastic, produced by elastic forces,

    • magnetic, produced by magnetic forces,

    • electrostatic, produced by Coulomb forces,

    • chemical, produced by Coulomb forces while chemical reactions occur (during chemical reactions atoms and electrons are rearranged, releasing or absorbing energy; this energy is electrical and manifests among nuclei, electrons, and molecules),

    • thermal, due to the kinetic energy of molecules and their relative position,

    • nuclear, caused by various nuclear forces (weak, strong).

    The potential energy variation due to displacement of a thermodynamic system in a gravitational field between elevations z1 and z2 is equal to the mechanical work of gravitational force. From Table 1.3, the expression of potential energy variation is written as

    Internal energy represents a summation of many microscopic forms of energy including vibrational, chemical, electrical, magnetic, surface, and thermal. The internal energy is proportional to the average force that molecules exert on the system boundary. Internal energy can be introduced based on the concept of ideal gas. According to the definition, ideal gas represents a special state of matter which can be delimited by a system boundary. Ideal gas is formed by a number (N) of freely moving molecules or atoms with perfect elastic behavior at collisions with each other and with the system boundary. It is assumed that:

    • all particles have rest mass (m > 0; the particles are not photons),

    • the number of particles with respect to the system volume is small,

    • the total volume of particles is negligible with respect to system volume,

    • collisions of particles with each other is much less probable than the collisions with the system boundary.

    Internal energy is an extensive quantity which plays the role of a state function. According to the thermodynamic definition a state function is the property of a system that depends only on current state parameters. When a change occurs, the state function is not influenced by the process in which the transformation is performed. Variation of internal energy can be expressed based on the specific internal energy which is defined by the relationship

    where m is the system mass; as inferred from the definition, the specific internal energy is an intensive property. Using the specific internal energy, the following expression for a system’s internal energy is obtained for a closed system

    The shape of the thermodynamic system can be arbitrary, but for simplicity we assume here a cubic volume as indicated in and the cube edge is l; it follows that the time between two collisions, approximated by the time needed for the particle to travel between two opposite walls, is

    Figure 1.3 A thermodynamic system filled with ideal gas.

    Based on the momentum conservation law, the force exerted by one particle on the wall, during collision, is

       (1.9)

    One also assumes that there is a uniform distribution of particle collisions for the three Cartesian directions; thus only one-third of particles exert force on a wall. Therefore, the pressure expression is determined by dividing the force (given by Equation (1.9)) with wall area A

       (1.10)

    where V = l A is the volume of the thermodynamic system and A is the surface area of the face, and

    is the gas density. The kinetic energy of a single gas particle can be expressed based on the average particle velocity; in this respect, Equation and it results in

       (1.11)

    The degree of freedom of monoatomic gas molecules is DOF = 3, because there are only three possible translation movements along Cartesian axes. According to its thermodynamic definition, temperature (T) is a measure of the average kinetic energy of molecules per degree of freedom. The quantitative relation between temperature and kinetic energy of one single molecule is according to

       (1.12)

    where kB is the Boltzmann constant defined in Table 1.2. Solving Equation (1.12) for T results in the thermodynamic expression for temperature, as follows:

       (1.13)

    Kinetic energy expression from Equation (1.11) can be introduced in Equation (1.13) and is obtained by

    where v is the molar specific volume

    and n is the amount of substance (number of moles). Therefore, the following thermodynamic definition of temperature is obtained, which is equivalent with that given in Equation (1.13)

       (1.14)

    (the universal gas constant) is given as follows:

    The value of the universal gas constant is given in which represents the ideal gas law, where v is the molar specific volume. It is possible to express the ideal gas equation on a mass basis if one introduces the gas constant

       (1.15)

    The ideal gas equation can be expressed using the gas constant R (in kJ/kmol K) and mass specific volume v (in m³/kg) as follows:

       (1.16)

    Here pressure, temperature, and specific volume represent a set of thermodynamic properties that completely define the thermodynamic state and the state of aggregation of a substance (e.g., solid, liquid, gas). The thermodynamic state of a system can be modified via various interactions, among which heat transfer is one. Heat can be added or removed from a system. When heat is added the internal energy of the system can increase, while the opposite happens when heat is removed. If heat is added the state of aggregation of the system passes from solid to liquid to gas.

    In Figure 1.4 the state diagram of water is exemplified. On the abscissa the heat addition to a thermodynamic system enclosing pure water is represented. On the ordinate the average temperature of the system is shown. For this diagram we assume that the internal energy of the system is zero at 250 K when the state of aggregation is that of subcooled ice. Ice reaches its melting point at 273.15 K. During the melting process, an ice-water mixture is formed. Due to phase change, the temperature remains constant (see figure), although heat is continuously added. At the end of the melting process all water is in a liquid state of aggregation. Water is further heated, and its temperature increases until it reaches the boiling point at 373.15 K. Additional heating produces boiling which evolves at constant temperature, while a water and steam mixture is formed. The boiling process is completed when all liquid water is transformed in steam. Further heating leads to temperature increase and generation of superheated steam.

    Figure 1.4 State diagram of water for 101.325 kPa.

    The pressure versus specific volume diagram of water is presented in Figure 1.5. This plot indicates the saturation lines where liquid and vapor reach the saturation temperature at a given pressure. Observe that the specific volume of saturated vapors is 1,000 times higher for normal boiling point isotherm. The normal boiling point isotherm corresponds to a temperature of 373.15 K. The specific volume along the boiling pallier can be expressed based on vapor quality, which is defined based vv′, v ″, the specific volumes of mixture, saturated liquid, and saturated vapor, respectively, according to

    Figure 1.5 Pressure versus specific volume diagram for water.

    where specific volumes are evaluated at the same pressure (or temperature). It can be observed that vapor quality varies between 0 and 1, with value 0 for saturated liquid and 1 for saturated vapor.

    Another useful diagram that shows the relationship between pressure, temperature, specific volume, and state of aggregation plots temperature versus specific volume. Figure 1.6 presents the Tv diagram for water, including the saturation lines, isobars, and constant vapor quality lines. Similar diagrams can be constructed for other substances.

    Figure 1.6 Temperature-specific volume diagram for water.

    The triple point represents that thermodynamic state where solid, liquid, and vapor can coexist. The triple point of water occurs at 273.16 K, 6.117 mbar, and the specific volume is 1.091 dm³/kg for ice, 1 dm³/kg for liquid water, and 206 m³/kg for vapor. Below the triple point isobar there is no liquid phase. A sublimation/desublimation process occurs which represents phase transition between solid and vapor.

    Between triple point isobar and critical point isobar three phases exist: solid, liquid, and vapor. In addition there are defined thermodynamic regions of subcooled liquid, two-phase, and superheated vapor. The subcooled liquid region exists between the critical isobar and the liquid saturation line (see Figure 1.6). The two-phase region is delimited by the liquid saturation line at the left, vapor saturation line at the right, and triple point isobar at the bottom. Superheated vapor exists above the vapor saturation line and below the critical isobar. At temperatures higher than the temperature of critical point and above the critical isobar there is a thermodynamic region denoted as "supercritical fluid region" where the substance is neither liquid nor gas, but has some common properties with gas and with liquid; supercritical fluids will be discussed in detail in other chapters of the book.

    The pressure versus temperature diagram is also an important tool that shows phase transitions of any substance. Figure 1.7 presents the PT diagram of water. There are four regions delimited in the diagrams: solid, vapor, liquid, and supercritical fluid. The phase transition lines are sublimation, solidification, boiling, critical isotherm, and critical isobar; the last two lines are represented only for the supercritical region (at pressure and temperature higher than critical).

    Figure 1.7 Pressure versus temperature diagram of water. Data from Haynes and Lide (2012).

    1.3 Equations of State and Ideal Gas Behavior

    An equation of state of a substance is a functional relationship which relates thermodynamic parameters and has the purpose of describing in a complete form the interrelation among state variables. For example, if one specifies the pressure and temperature one must be able to calculate with the help of an equation of state the specific volume and the specific internal energy. If these quantities can be determined, any other thermodynamic parameter can be calculated using standard thermodynamic relations.

    According to the so-called state postulate a simple compressible system is completely specified by two independent, intensive properties. Usually the following thermodynamic properties are taken as independent variables: pressure, specific volume, and temperature. Once the independent variables are specified all other thermodynamic parameters can be derived for any particular substance and any given thermodynamic state. There are two types of equations of state:

    • thermal equation of state—as a function of P, v, T

    • caloric equation of state—as a function of u, v, T, where u represents the specific internal energy

    The ideal gas law—expressed by Equation (1.16)—is a thermal equation of state because it relates pressure, specific volume, and temperature. The caloric equation of the state of ideal gas can be derived from Equation (1.11) which represents the kinetic energy of a single molecule. Accordingly, for 1 mol the internal energy is N, the caloric equation of state for ideal gas is

    . The ideal gas model is valid only for low pressure and large specific volume. The model becomes inexact at high pressure and temperature in the vicinity of critical point, when gas becomes much dense and is influenced by the vapor–liquid transition. In addition, the gas compressibility decreases greatly close to saturation line and critical point. In order to account for compressibility of a substance, the compressibility factor Z is introduced, which is a measure of alleviation of actual specific volume using the prediction of ideal gas law for the same pressure and temperature conditions. Compressibility factor thus expresses the deviation of a substance from ideal gas behavior as defined below:

    where specific volume is expressed on mass basis.

    The order of magnitude is about 0.2 for many fluids. For accurate thermodynamic calculations compressibility charts can be used, which express compressibility factor as a function of pressure and temperature. In this way, an equation of state obtained based on compressibility factor is written as follows:

    where the compressibility factor is a function of pressure and temperature.

    According to the principle of corresponding states, the compressibility factor has a quantitative similarity for all gases when it is plotted against reduced pressure and reduced temperature. The reduced pressure is defined by the actual pressure divided by the pressure of the critical point

    where subscript c refers to critical properties and subscript r to reduced properties. Analogously, the reduced temperature is defined by

    Compressibility charts showing the dependence of compressibility factor on reduced pressure and temperature can be obtained from accurate PvT data for fluids. These data are obtained primarily based on measurements. Accurate equations of state exist for many fluids; these equations are normally fitted to the experimental data to maximize the prediction accuracy. A generalized compressibility chart

    is presented in Figure 1.8; the chart was obtained with Engineering Equation Solver software (EES, 2013), with which compressibility factor has been calculated for thirteen fluids and averaged.

    Figure 1.8 Generalized compressibility chart obtained for 13 fluids [Engineering Equation Solver software is used for Z prediction with real gas equations of state].

    There are many equations of state published in the literature. One of the most basic equations of state is that of Van der Waals, which is capable of predicting the vapor and liquid saturation line and a qualitatively correct fluid behavior in the vicinity of the critical point. This equation is described by the following expressions

    where subscript r indicates reduced quantities. The reduced volume and the reduced internal energy are defined by

    In thermodynamics, there are some special processes—of practical importance—during which P, or v, or T remains constant. If there is no heat exchange with the exterior (dq = 0) then the process is called adiabatic. If a process is neither adiabatic nor isothermal it can be modeled as polytropic. Table 1.4 summarizes the principal features of simple processes for ideal gas. Figure 1.9 presents a P – v diagram for four simple processes with ideal air, modeled as ideal gas.

    Table 1.4

    Simple Thermodynamic Processes and Corresponding Equations for Ideal Gas Model

    Figure 1.9 Simple processes with ideal gas represented in P  –  v diagram.

    1.4 Laws of Thermodynamics

    In thermodynamics there are two main laws (first and second law) which describe the behavior of thermodynamic systems regardless of their type. There is also a third law of thermodynamics—commonly called the zeroth law which refers to the state of thermodynamic equilibrium. In general, thermodynamic analysis assumes that thermodynamic systems are at equilibrium, although for some cases non-equilibrium thermodynamics methods may be the better choice for analysis. Two thermodynamic systems are said to be in thermal equilibrium if they cannot exchange heat, or, in other words, they have the same temperature. Two thermodynamic systems are in mechanical equilibrium if they cannot exchange energy in the form of work. Two thermodynamic systems are in chemical equilibrium if they do not change their chemical composition. An insulated thermodynamic system is said to be in thermodynamic equilibrium when no mass, heat, work, chemical energy, etc. is exchanged between any parts within the system.

    The zeroth law of thermodynamics is a statement about thermodynamic equilibrium expressed as follows: if two thermodynamic systems are in thermal equilibrium with a third, they are also in thermal equilibrium with each other. A system at internal equilibrium has a uniform pressure, temperature, and chemical potential throughout its volume.

    Energy is neither created nor destroyed. This is one of the statements of the First Law of Thermodynamics (FLT) which is known as the energy conservation principle. The FLT can be phrased as you can’t get something from nothing. If one denotes E the energy (in kJ) and ΔEsys the change of energy of the system, then the FLT for a closed system undergoing any kind of process is written in the manner illustrated in Figure 1.10. There are two mathematical forms for FLT, namely on an amount basis or on a rate basis. These mathematical formulations are indicated by

    Figure 1.10 Explanatory sketch for the first law of thermodynamics.

    Energy can be transferred to or from a thermodynamic system in three basic forms, namely as work, heat, or through energy associated with a mass crossing the system boundary. Therefore one has

    and the variation of system energy—assuming that the mass of the closed system is m—is

    where index 1 represents the initial state and index 2 the final state and e is the specific total energy of the system, comprising internal energy, kinetic energy, and potential energy, expressed as follows:

    In classical thermodynamics there is, however, a sign convention for work and heat transfer which is the following:

    • the heat is positive when given to the system, that is Q = ∑ Qin − ∑ Qout is positive when there is net heat provided to the system.

    • the net work, W = ∑ Wout − ∑ Win is positive when work is generated by the system.

    Using the sign convention, the FLT for closed systems becomes

    The above equation can be expressed on mass specific basis as follows:

    where Q = m q and W = m w and ΔEsys = m (e2 − e1). Furthermore, FLT can be written in differential form as follows:

    Taking into account that for a closed system dw = P dv, it results that

    With the assumption that the changes in kinetic and potential energies are negligible, the FLT for a closed system becomes

       (1.17)

    If the system is a control volume, then the energy term will comprises the additional term of flow work. In this case the total specific energy of a flowing matter is

    The term u + P v from above equation represents a state function named enthalpy (or flow energy), denoted with h

    Using enthalpy formulation, the FLT for a control volume that has neither velocity nor elevation is d(h − Pv) = dq − Pdv which can be expanded to the following equation: dh − Pdv − vdP = dq − Pdv. Thus, the FLT for control volumes (open systems) becomes

       (1.18)

    In addition, one remarks that the total specific energy of a flowing matter can be written with the help of enthalpy

       (1.19)

    The SLT for control volume, using the sign convention for heat and work, is formulated mathematically, in rate form, in the following way:

      

    (1.20)

    An important consequence of the FLT is that the internal energy change resulting from any process is independent of the thermodynamic path followed by the system during thermodynamic transformations. In turn, the rate at which the internal energy content of the system changes is dependent only on the rates at which heat is added and work is done. Observe from Equation (1.17) that internal energy is a function of specific volume; recall that internal energy is also a function of temperature. Therefore one has u = u(Tv) and, accounting for Equation (1.17), the total derivative of internal energy is

    The above equation defines the specific heat at constant volume Cv, which represents the amount of heat required to increase the temperature of a constant volume system by 1° K; as is observed from above equation, du = dq = Cv dT. The specific heat is equal to the change in internal energy with temperature at constant volume

       (1.21)

    In addition it results in the following thermodynamic identity

    Similarly, from Equation (1.18), enthalpy is a total derivative of temperature and pressure, h = h(TP). The total differential form of enthalpy becomes

    an expression that defines the specific heat at constant pressure Cp, which represents the amount of heat required to increase the temperature of a system evolving at constant pressure at 1° K. The specific heat is the change of enthalpy with temperature at a constant pressure defined according to

       (1.22)

    In addition it is observed from above equation that for an isobaric process the heat exchange is governed by the equation dh = dq = CpdT, and the following thermodynamic identity is noted:

    For an ideal gas it is remarked that dh = d(u + P v) = du + R dT, or by replacing dh and du with dh = Cp dT and du = Cv dT one obtains the well-known Robert Meyer equation

    For ideal gas it was demonstrated above that kinetic energy associated with molecules is the same as its internal energy, u = 1.5 R T. It can be remarked that for ideal gas the internal energy is a function of temperature only. Therefore, specific heat for ideal gas is Cv = 1.5 R and Cp = 2.5 R. For ideal gases, internal energy is a function of temperature only, so du = cv dT. From the definition h ≡ u + P v it can be inferred that the enthalpy of ideal gas (P v = R T) is dependent enthalpy only because h(T) = u(T) + R T. In conclusion one has dh = Cp dT. In addition, from enthalpy definition it results that dh = du + R dT.

    The ratio of specific heat at constant pressure and constant volume is known as the adiabatic exponent

    which for monoatomic ideal gas has the value 1.4, and a value of 5/3 = 1.67 for diatomic gas.

    Observe that the specific heat of ideal gas is constant. In some literature sources an ideal gas with constant specific heat is denoted as "perfect gas," whereas gases that behave according to ideal gas law (P v = R T) but show specific heat variation with temperature are denoted as "ideal gases."

    Another thermodynamic law is the second law of thermodynamics, which provides a way to predict the direction of any process in time, to establish conditions of equilibrium, to determine the maximum attainable performance of machines and processes, and to assess quantitatively the irreversibilities and determine their magnitude for the purpose of identifying ways of improvement of processes. Also SLT is the fundamental law based on which the absolute scale of temperature is defined. Furthermore, SLT is useful in evaluating various relevant thermodynamic properties based on experimental data.

    In order to introduce SLT the concepts of reversibility and irreversibility must be discussed. It is known that a thermodynamic process is reversible if during a transformation both the thermodynamic system and its surroundings can be returned to their initial states. All infinitesimal changes through which the process evolves must be reversible, such that the overall process is reversible. A process performed by a thermodynamic system is irreversible if it cannot return to the initial state because of dissipation that occurs (irreversibly) by friction, heat rejection, electrical, chemical, or mechanical losses etc. In any real application there is dissipation; therefore, reversibility is only an idealization that is important for theoretical point of view. Reversible processes can be characterized as follows:

    • Externally reversible—a process with no associated irreversibilities outside the system boundary.

    • Internally reversible—a process with no irreversibilites within the boundary of the system during the process.

    • Totally reversible—a process with no irreversibilities within the system or surroundings.

    The classical statements of SLT which say in essence that heat cannot be completely converted into work, although the opposite is possible (a form of work can be completely converted into any other form of work) are as follows.

    • The Kelvin-Plank statement. It is impossible to construct a device, operating in a cycle (e.g., a heat engine), that accomplishes only the extraction of heat energy from some source and its complete conversion to work. This simply shows the impossibility of having a heat engine with a thermal efficiency of 100%.

    • The Clausius statement. It is impossible to construct a device, operating in a cycle (e.g., a refrigerator and heat pump), that transfers heat from the low-temperature side (cooler) to the high-temperature side (hotter).

    The Clausius inequality provides a mathematical statement of the SLT, namely

       (1.23)

    where the circular integral indicates that the process must be cyclical.

    At the limit when the inequality becomes zero then the process is reversible (ideal situation). A useful mathematical artifice is to attribute to the integral from Equation (1.23) a new physical quantity. This will define entropy generated by the system Sgen

       (1.24)

    Any real process must have positive generated entropy; the following cases may thus occur: (i) Sgen > 0, real, irreversible process; (ii) Sgen = 0, ideal, reversible process; (iii) Sgen < 0, impossible process. Therefore, the mathematical formulation of the SLT is

    Entropy is an extensive property. Generated entropy of a system during a process is a superimposition of entropy change of the thermodynamic system and the entropy change of the surroundings

       (1.25)

    For a reversible process Sgen = 0, Equation (1.25) shows that entropy change of the system is the opposite of the entropy change of the surroundings:

    For an irreversible process the following inequality must be true

    which is explained by the existence of generated entropy that is always positive.

    Although the change in entropy of the system and its surroundings may individually increase, decrease, or remain constant, the total entropy change (the sum of entropy change of the system and the surroundings or the total entropy generation) cannot be less than zero for any process. Note that entropy change along a process 1–2 is defined based on the integral

    therefore by differentiation one obtains

       (1.26)

    Entropy quantifies the molecular random motion within a thermodynamic system and is related to the thermodynamic probability (p) of possible microscopic states as indicated by the Boltzmann equation

    The engineering usefulness of SLT consists of the ability to define and predict the performance limits of systems. As is known, a heat engine is a device operating based on a specific thermodynamic cycle, which eventually generates useful work. In the reverse process, a heat pump consumes work in order to increase the temperature of a working fluid and generate heating. A refrigerator is similar to a heat engine, except that it generates cooling. The environment plays the role of a large thermal energy reservoir at a specified temperature. A thermal energy reservoir is a thermodynamic system with relatively large thermal energy capacity that can absorb or deliver large quantities of heat without observable changes in its temperature.

    A heat source represents a thermal reservoir capable of providing thermal energy to other systems. A heat sink represents a thermal reservoir capable of absorbing heat from other systems. The environment can play roles of both heat source or heat sink, depending on the application. A heat engine operates cyclically by transferring heat from a heat source to a heat sink. While receiving more heat from the source (QH) and rejecting less to the sink (QC), a heat engine can generate work (W). As stated by the FLT, energy is conserved, thus QH = QC + W.

    A typical black box representation of a heat engine is presented in Figure 1.11a. According to the SLT the work generated must be strictly smaller than the heat input, W < QH. The thermal efficiency of a heat engine—also known as energy efficiency—is defined as the net work generated by the total heat input. Using notations from Figure 1.11a, energy efficiency of a heat engine is expressed (by definition) with

       (1.27)

    Figure 1.11 Conceptual representation of a heat engine (a) and heat pump (b).

    If a thermodynamic cycle operates as a refrigerator or heat pump, then its performance can be assessed by the Coefficient of Performance (COP), defined as useful heat generated per work consumed. As observed in Figure 1.11b, the energy balance equation (EBE) for a heat pump is written as QC + W = QH, according to SLT QH ≥ W (this means that work can be integrally converted in heat). Based on its definition, the coefficient of performance is

       (1.28)

    The Carnot cycle is a fundamental model in thermodynamics representing a heat engine (or heat pump) that operates between a heat source and a heat sink, both of them being at constant temperature. This cycle is a conceptual (theoretical) cycle and was proposed by Sadi Carnot in 1824. The cycle comprises fully reversible processes, namely two adiabatic and two isothermal processes. Figure 1.12 illustrates two Carnot cycles for air, modeled as ideal gas: (a) is a power cycle, and (b) is a reversed (refrigeration or heat pump) cycle. The efficiency of the Carnot cycle is independent of the working fluid which performs the cyclical process.

    Enjoying the preview?
    Page 1 of 1