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

Only $11.99/month after trial. Cancel anytime.

Thermodynamics: Foundations and Applications
Thermodynamics: Foundations and Applications
Thermodynamics: Foundations and Applications
Ebook1,733 pages18 hours

Thermodynamics: Foundations and Applications

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Designed for both undergraduate and graduate students, this authoritative milestone in the foundational development of thermodynamics provides a unique reference for all physicists and engineers. Basic concepts and applications are discussed in complete detail with attention to generality and logical consistency, removing ambiguities and limitations of traditional presentations. Worked-out examples and end-of-chapter problems illustrate the use of energy and entropy balances as powerful analytical keystones in physics and engineering.
The text provides material for undergraduate and graduate courses. At the introductory level, it covers heat engines, stable-equilibrium-state models for ideal-gas, incompressible-fluid and solid behaviors, heat, work and bulk-flow interactions, thermodynamic efficiency, energy conversion systems, energy, and availability/ At the intermediate level, it covers ideal and nonideal mixtures, chemical reactions, chemical equilibrium, and combustion.
At the advanced level, the unique non-traditional order of exposition of the basic concepts and principles (system, property, state, process, first law, energy, equilibrium, stable equilibrium, second law, entropy) allows rigorous general definitions of energy and entropy valid for all systems (large and small, few- and many- particles) and all states (stable and non-stable equilibrium, as well as non-equilibrium). In particular, entropy is defined before and independently of the definitions of temperature and heat, and of the simple-system model for many-particle systems.
LanguageEnglish
Release dateJul 12, 2012
ISBN9780486135182
Thermodynamics: Foundations and Applications

Related to Thermodynamics

Related ebooks

Mechanical Engineering For You

View More

Related articles

Reviews for Thermodynamics

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Thermodynamics - Elias P. Gyftopoulos

    Applications

    1 How to Study with This Book

    In thermodynamics, we model our experiences of physical phenomena in terms of concepts, mathematical symbols that represent the concepts, and relations between symbols that represent interdependences between concepts. Some aspects of this procedure are readily understandable even by people without much expertise in the subject. Other aspects are somewhat abstract and require systematic study. In any case, the results are fascinating and very practical. We use models to describe physical phenomena at one instant of time and to make predictions as a function of time. For example, we may wish to evaluate the humidity of the atmosphere on a hot summer day, or design and control the operation of a refinery that transforms crude oil into petroleum products.

    In this chapter we describe some salient features of the models of physical phenomena without being concerned about the precise meanings of the terms we use. We do this especially for readers being exposed to the subject for the first time. We jump in the middle of energy and entropy balances, in much the same way that in a fluid mechanics text one jumps in the middle of mass and momentum balances, or in a chemistry text one jumps in the middle of stoichiometric balances of atomic nuclei.

    After seeing how these balances play a central role in solving a few sample problems, we hope that readers will be sufficiently motivated to pursue a systematic study of the applications and a deeper scrutiny of the foundations of thermodynamics. This pursuit can be approached along four different paths: (1) an undergraduate introductory path, (2) an advanced undergraduate path with emphasis on applications, (3) a graduate path with emphasis on the general foundations, and (4) a graduate path with emphasis on advanced topics. Each path consists of different parts of the book. We describe the parts that belong to each path at the end of this chapter.

    1.1 Modeling Physical Phenomena

    Some concepts are essential in all models of physical phenomena. They are the basic tools with which scientists and engineers reason and report their findings. Among them are the concepts of system, property, and the law of motion or change with time. Any model of any physical phenomenon includes the definitions of a system—the object of study, its properties—the characteristics of the system that can be measured at one instant of time by means of well-defined operations, and the law of motion—the relation between values of properties at different instants of time.

    The requirement that properties must be measurable provides the link between the model and the phenomena it describes. In some instances, this link is both readily understandable and fairly direct. In others, the link may be conceptually complicated and operationally indirect. For example, the position of a particle of a one-particle system is a property that we can easily visualize and readily measure. In contrast, the acceleration of a particle is harder to picture, and its measurement more involved. Even more complicated is the idea of the mass of the particle and its measurement. Although some properties are more abstract than others, they are all measurable and are all linked to the phenomena they describe.

    The law of motion is the necessary tool for making predictions as a function of time. It helps us not only to make forecasts about natural occurrences, but also to explore natural phenomena for the benefit of our society. For example, we use the law of motion to determine the location of a celestial body a year hence. Again, we use the law of motion to design and control the flight of an airplane from one city to another. Some models may lack a law of motion but still be very valuable because they define either specific interrelations among different properties at one instant of time, or specific facets of the law of motion, such as the time dependence of only a few of the properties, or both.

    Over the centuries, a large number of different models have been developed, each for a particular class of physical phenomena. Despite the differences, however, certain properties have emerged as common features of all models, and provide an excellent degree of unity in our understanding of the workings of nature. Of course, the very ambitious and worthy undertaking of a unified interpretation of our experiences by means of an all-encompassing model must continue. In all likelihood, such an undertaking will remain forever incomplete because we will never be able to assert that we have observed all physical phenomena.

    Even if the all-encompassing model were known, we would not use it in practice to solve every specific scientific problem, and every specific engineering problem because such an approach would probably be very complicated. Instead, we adapt our consideration to the problem at hand and introduce simplifications and approximations. In doing so, however, we make sure that as many of the common features of all models are retained as practicable, and that these features satisfy all the conditions that are widely accepted as valid. For example, from the various models that have been developed to date, a firm conclusion is that energy is neither created nor destroyed. So, in any modeling of a problem, we make sure that energy is a property and that energy is conserved.

    Some of the well-established general features of all models and the way they influence our approach to problem solving are discussed immediately after we introduce the idea of accounting and balances in the next section.

    1.2 Accounting and Balances

    The procedure of performing a balance of some quantity of interest is part of our everyday experience. Outside the domain of natural phenomena, the most familiar example is what we call accounting, whether applied to our personal bank savings or to the finances of a business. Although this example belongs to the domain of economics, we discuss it here because it brings out certain self-evident features of the properties that must be balanced and that are also features of energy and entropy.

    To make these ideas specific, we consider a safe in a bank in which a family keeps its savings in the form of golden coins. We call these savings account A, denote the number of coins in the safe by NA, and recognize that it has three useful and practical features. The first feature is that NA can be measured at any instant of time by a simple procedure—counting. At time t1, we denote the value of NA by N1A so that if the number of coins is 2000, we write N1A = 2000 coins. It is evident that NA must play a key role in any accounting model.

    The second feature that makes the quantity NA useful in accounting is that it is additive. This means that if NC denotes the number of coins in a composite account C consisting of two accounts A and B with numbers of coins denoted by NA and NB, respectively, then at any instant of time t1 the value N1C of NC equals the sum of the values N1A and N1B, that is, N1C = N1A + N1B. The feature of additivity is useful because it allows us to keep track of a financial situation even if it consists of an intricate structure of different accounts.

    The third feature is that different accounts may be made to interact with each other, that is, coins may be transferred from one account to another. If the interacting accounts are A and B, we denote the coins transferred from account B into account A by the symbol NA←, where we use the arrow in the superscript because an amount transferred, like a vectorial quantity, is defined not only by a numerical value—the number of coins transferred—but also by the direction of the transfer. Thus NA← = 1000 coins means that 1000 coins are transferred from account B to account A, whereas NA→ = 1000 coins means that 1000 coins are transferred from account A to account B. Because the arrow points either into A or out of A, we can use one symbol to denote both transfers, and either positive or negative values to specify whether the flow is actually into or out of A. For example, if we choose the symbol NA← we assign a positive value when coins are transferred into A, and a negative value when coins are taken out of A. Clearly, if NA← = 2000 coins, then NA→ = —2000 coins.

    Performing a balance on the number of coins NA of account A over a period of time from time tl to time t2 means that we evaluate the change in the value of NA during that time interval, that is, the difference N2A — N1A, and compare it with the net number of coins N12A← that were transferred into the account as a result of all the transactions that occurred between t1 and t2. For example, account A might have had N1A = 4000 coins at a time t1, N2A = 3000 coins at time t2, and have experienced two transactions, NA← = —2500 coins, and NA←= 1500 coins, so that N12A← = —2500 + 1500 = —1000 coins over the period from t1 to t2.

    Comparison of the values N2A — N1A and N12A← brings out important features that have to do with the nature of the characteristic property N, and the nature of the dynamics of the system—the account. If for any time interval and any transactions, we find that the balance

    is satisfied, then we say that N is conserved. Of course, we know from experience that the number of coins in a well-guarded safe is conserved. Nevertheless, this obvious conservation principle—self-evident identity—is used by accountants to make sure that all accounts are in good order.

    The number of coins is not the only property of interest of each account A. Another characteristic of even greater importance is the purchasing power of the coins. We denote it by PA. At any instant of time t1 the purchasing power of the coins, P1A, represents the fraction of the gross national product (GNP) that at time t1 can be bought with the coins in the safe. It is measured in dollars per dollar of GNP or some other currency, and has the same three basic features as the number of coins, that is, measurability, additivity, and transferability.

    Over a period of time from t1 to t2, the purchasing power of the coins in a safe can be altered by two factors: the transfer of purchasing power P12A← from other accounts as a result of transfers of coins into and out of the safe; and the creation or destruction of purchasing power as a result of phenomena such as deflation or inflation of prices, and variations in the exchange rate of golden coins to currency. Denoting the latter effects by Pi, we can relate the purchasing power P2A at time t2 to the purchasing power P1A at time tl by the balance

    If gold appreciates as fast as prices are inflated over the period from t1 to t2, then Pi is zero and the purchasing power is affected only by transfer transactions. On the other hand, if gold does not appreciate as fast as prices are inflated, Pi < 0, and the purchasing power of a dormant account (P12A← = 0) is being destroyed even though the number of coins in the account is conserved.

    In conclusion, our simple accounting example shows that every golden-coin account A has the property NA, representing the number of coins in the account, and the property PA, representing the purchasing power of the coins in the account. Both properties are always defined at any instant in time and in all circumstances, and both are measurable, additive, and transferable. These features make both property NA and property PA very useful because we can perform balances on them over a time period of interest. Moveover, whereas property NA is conserved, that is, its balance equation has no creation or destruction term [equation (1.1)], in general, property PA is not conserved, that is, its balance equation contains a creation or destruction term Pi [equation (1.2)].

    In the domain of natural phenomena, two properties play a similarly central role in any well-defined model: energy and entropy. Similarly to the number of coins NA, energy is a conserved property. Similarly to the purchasing power PA, in general, entropy is not a conserved property. We discuss these features of energy and entropy in the next section.

    1.3 Energy and Entropy Balances

    Most of this book is devoted to defining, understanding, and using energy and entropy in solving practical problems. Here we give a preview of how energy and entropy play such a central role in modeling the behavior of all systems under all conditions.

    Both energy and entropy have the three basic features of measurability, additivity, and transferability that we discuss in the preceding section. The system that we define to model any physical phenomenon has both energy and entropy as properties. Each of these properties can be measured at any instant of time by means of well-defined procedures. We denote energy by E, entropy by S, and the values of E and S at time t1 by E1 and S1 respectively. For example, upon applying the measuring procedures that define energy and entropy to a given system at time t1 we could find E1 = 400 J and S1 = 5 J/K, where J stands for the units of measure of energy, and J/K for the units of measure of entropy. In due course we will see that one set of units are joules for energy and joules per kelvin for entropy. Because these measuring procedures are applicable to any well-defined system in any condition and at any instant of time, we say that energy and entropy are general properties of any model of any physical phenomenon.

    Each of the properties E and S is additive. If a composite system C consists of two subsystems A and B, and we denote the energies and entropies of the three systems by EC, EA, EB and SC, SA, SB, respectively, then at any instant of time t1 the value E1C equals the sum of the values of E1A and E1B, that is, E1C = E1A + E1B, and the value of S1C equals the sum of the values of S1A and S1B, that is, S1C = S1A + S1B. The additivity feature is useful because it allows us to keep track of the values of energy and entropy even in complicated systems consisting of intricate structures of subsystems, such as chemical industrial plants, power plants, or laboratory experiments.

    Two systems can interact with each other and, as a result of the interaction, exchange energy and entropy. If the interacting systems are A and B, we denote the amounts of energy and entropy transferred from system B to system A by EA←and SA← respectively, where we use the arrow in the superscript because an amount transferred is defined not only by the numerical value—the amount of energy or entropy transferred— but also by the direction of the transfer. As we did for the accounting example, we can denote a transfer of 200 J of energy and 2 J/K of entropy into system A by writing either EA←= 200 J and SA← = — 2 J/K or, equivalently, EA→= -200 J and SA→ = -2 J/K.

    The three features of energy and entropy just listed—measurability, additivity, and transferability—make these two properties suitable for setting up balances in much the same way that we do in the accounting example of the preceding section. Performing balances of energy and entropy for a system A over a period of time from t1 to t2 means that we evaluate the changes E1A E1A in energy and S1A S1A in entropy and compare them with the respective net amounts of energy E12A← and entropy S12A← transferred into the system as a result of all the interactions experienced by A during the time period between t1 and t2.

    Comparison of the values of E2A — E1A and E12A← brings out the important feature that energy is conserved, that is, that the energy balance is always of the form

    whether t2 > t1 or t1 > t2.

    Comparison of the values S2A — S1A and S12A← brings out the important feature that entropy cannot be destroyed but can be created inside the system as time goes on. If t2 > t1, we express this feature by writing the entropy balance in the form

    where SAirr is such that

    that is, SirrA is nonnegative. Equation (1.4) may also be written in the form

    In due course, we will call SirrA the entropy created or generated by irreversibility inside system A.

    The energy and entropy balances, equations (1.3) and (1.4) [or (1.3) and (1.6)], are the cornerstones of the dynamical analysis of all physical phenomena. Yet they are among the most abstract concepts of the description of nature. Perusing once more the analogy with the accounting example in the preceding section, we observe that the value of the energy EA of any system A can change only as a result of interactions that transfer energy into or out of the system, in much the same way that the value of the number of coins NA in an account A can change only as a result of transactions that transfer coins into or out of the account. In contrast, the value of the entropy SA of any system A can change even in the absence of any interaction, that is, spontaneously, because of the workings of nature, in much the same way that the value of the purchasing power PA can change even in the absence of any banking transactions because of the workings of the economy. In fact, we will see that the relation between energy and entropy has several analogies to that between the number of golden coins and their purchasing power.

    We never use the accounting example again, but we devote several chapters of this book to defining the meanings of energy and entropy, and explore their interrelations and far-reaching implications on the behavior of various systems. A better understanding of the interrelations and implications contributes greatly to our ability to design better equipment to control various physical phenomena.

    1.4 Applications of Energy and Entropy Balances

    The problem of understanding the physical meanings of energy and entropy can be tackled at various levels of depth and from different standpoints. The most basic approach would be in terms of an advanced physics course based on a general and all-encompassing model of physical phenomena. Such a course would require a good background in mathematics and a strong interest in abstract thinking. The subject is fascinating, the open questions are challenging, and the horizon for scientific contributions is exciting. But such an approach would be too basic and would hardly convey the usefulness of thermodynamics in analyses of applied physics and engineering problems.

    The approach we take in this book is to start from a nonmathematical summary of concepts and statements that capture only the most general features of the all- encompassing model known to date, and to use these concepts and statements in the study of a number of broad classes of problems. The study of this approach may be pursued along four alternative paths, each providing material for a regular one-semester course, and each being outlined in Section 1.6.

    In an introductory course, the meaning of energy and entropy can be grasped gradually and indirectly through use of these two properties in solving practical problems, and in describing the behavior of physical systems in what are called the thermodynamic equilibrium or stable equilibrium states. For such states, energy and entropy can be related to other properties, such as the equilibrium properties temperature and pressure, which in turn are linked to the object of study by familiar experimental instruments, such as a thermometer and a pressure gauge. For stable equilibrium states, the approach can be viewed as proceeding along two parallel but interrelated lines of analysis.

    The first line of analysis focuses attention on relating energy and entropy differences to more readily measurable properties by means of models of the stable equilibrium behavior of matter in selected ranges of conditions. Examples of such models are the perfect-gas model, and the perfect-incompressible model for solids or fluids. This line of analysis addresses the left-hand sides of the energy and entropy balances [equations (1.3) and (1.4)]. For example, for stable equilibrium states of a system A that behaves according to the perfect-gas model, energy and entropy differences can be expressed as

    where CVA, CpA, and RA are constants, and T2A, T1A and p2A, p1A are, respectively, the temperatures and the pressures of system A at times t2 and t1 Again, for stable equilibrium states of a system B that behaves according to the perfect-incompressible- fluid model, energy and entropy differences can be expressed as

    where CB is a constant, and T2B and T1B are the temperatures of the system at times t2 and t1 respectively. In due course, we devote several chapters to defining these and other models.

    The second line of analysis focuses attention on evaluating the exchanges of amounts of energy and entropy associated with selected types of interactions between systems. Examples of such types of interactions are work, heat, and bulk flow. This line of analysis addresses the right-hand sides of the energy and entropy balances. For example, for a work interaction experienced by a system A, energy and entropy transfers can be expressed as

    where the energy transfer W←is called Work, Again, for a heat interaction experienced by a system A, we have

    where the energy transfer Q← is called heat, and TQ is the temperature at the boundary of system A where the interaction occurs. In due course, we devote an entire chapter to defining work and heat interactions. Still another type of interaction is bulk flow, for which we have

    where mA← is the amount of mass transferred into system A as a result of the interaction, and h, s, ξ, g, and z are properties of the transferred material that are defined in due course. If more than one of these interactions occurs, the energy and entropy transfers are sums of the respective amounts for the relevant interactions.

    1.5 Problem Solving

    The energy and entropy balances, and the two lines of analysis just cited, constitute the primary tools for the systematic approach to solving the applied physics and engineering problems discussed in this book. The approach may be subdivided into four steps.

    In the first step, we define the system—the object of study. The step is important because it determines the ease of the solution. The selection of the system requires experience and a careful examination of the statement and the questions of the problem. For each question we may have to define a system that includes as many of the physical objects of the problem as possible, and that experiences at its boundary one or more of the interactions that must be evaluated.

    In the second step, we model the system. Usually, we look inside the boundary of the system and partition it into a collection of subsystems for each of which we can evaluate the changes in energy and entropy. For example, we might select each subsystem so as to be able to model its behavior as that of a perfect gas, perfect incompressible fluid, or cyclic device, that is, a device that experiences zero net changes of all its properties. If successful, this step allows us to express energy and entropy differences in terms of more readily measurable properties such as temperatures and pressures. Some of the latter properties may be known; others may be unknown.

    In the third step, we model the interactions occurring at the boundary of the system in terms of the transfers of energy and entropy that they contribute. For example, we may conclude that work, heat, and bulk-flow interactions are adequate to describe all the transfers of properties at the boundary of the system. The purpose of this step is to express the transfer terms E12A← and S12A← in the right-hand sides of the energy and entropy balances as sums of transfer terms, each referring to a specific interaction. Again, some of these transfer terms may be known, others may be unknown.

    In the fourth step, we combine the results of the second and third steps and form the energy and entropy balances. The energy balance equation can be solved for one of the unknowns of the problem. The entropy balance can be solved for SirrA if all the other variables are known, or it can be solved for another unknown if SirrA is given as part of the problem. For example, a question might imply that SirrA = 0.

    The foregoing four-step problem-solving scheme is succinct and captures only some of the essential ingredients of analyses based on energy and entropy balances. Proceeding along one of the four paths outlined in Section 1.6, the reader will find that the ideas so quickly summarized here are gradually clarified and, in a sense, guide the flow of the material.

    We conclude this section with an example representative of the kind of practical questions that can be answered by the use of energy and entropy balances.

    Example 1.1.  A block of metal, B, is immersed in a pool of liquid water, L. The block and the water interact with each other and nothing else. Temperature measurements on the block and on the liquid yield the values T1B and T1L at time t1 and only the value T2B at time t2. The engineer who performed the measurements forgot to tell us whether time t1 was before or after t2. Can we find the value of T2L , and infer the time sequence from the measured values of the temperatures?

    Solution.

    Step 1.   To answer the question we define as our system the composite A consisting of the block of metal B and the liquid water L.

    Step 2.   We model block B as a perfect incompressible solid, and the liquid water L as a perfect incompressible liquid, each with energy and entropy differences given by equations (1.9) and (1.10). Using these equations and the fact that energy and entropy are additive properties, we can express the terms E2A — E1A and S2A — S1Aon the left-hand sides of the energy and entropy balances [equations (1.3) and (1.4)] in terms of T1B, T1L, T2B , and T2L, so that

    Step 3.   No interactions occur at the boundary of system A because the block and the liquid water do not interact with anything else than each other. Thus no energy and no entropy are transferred into or out of the system A, that is, E12A← = 0 and S12A = 0.

    Step 4.   Substituting equations (1.17) and (1.18) in the left-hand sides of the energy and entropy balances [equations (1.3) and (1.4)], and using E12A← = 0 and S12A = 0 in the right-hand sides, we find that

    Upon solving equation (1.19), we determine T2L in terms of the measured temperatures and the two constants CB and CL. Substituting T2L in relation (1.20), we conclude that if

    then t2 > t1. On the other hand, if

    then t2 < t1.

    1.6 Four Courses of Study

    The topics included in this book are not meant to be covered either by a single course or by a sequence of interrelated courses. Different parts may be combined as one-semester offerings. Each offering serves different programmatic and student needs.

    For example, an undergraduate course for engineering students may begin with the ideas that energy and entropy are two properties of a system. Rather than through the formal definitions of these properties, the student is assisted to grasp their meanings and significance by repeated applications. Chapter 14 is recommended as the starting point of such courses. It summarizes the essential concepts without emphasis on subtleties, and without compromising and misleading arguments. Applications are selected from Chapters 15 to 31 as discussed in the following suggested outlines.

    1. Undergraduate Course: Introductory Thermodynamics. This course is for sophomores or juniors in the schools of science and engineering. From the end of the present chapter, jump directly to Chapter 14. Study Chapters 14 to 25, except for Sections 17.6 to 17.8, 20.5 and 20.6, and 21.3 to 21.6.

    Heat engines, heat pumps, and refrigeration units are discussed in Chapter 15; inter-relations among stable-equilibrium-state properties of systems with volume as the only parameter in Chapter 16; the simple-system model in Chapter 17; properties of pure substances, and models for ideal-gas and incompressible-fluid behaviors in Chapters 19 to 21; bulk-flow interactions in Chapter 22; devices used in energy-conversion and materials-processing systems in Chapter 23; optimum performance and thermodynamic efficiency in Chapter 24; and energy-conversion systems in Chapter 25.

    2. Undergraduate Course: Advanced Topics in Thermodynamics. If a program foresees a second semester on thermodynamics, complete the study of Chapters 16 and 17; and study Chapters 26, 27, and 29 to 31, except for Sections 27.5 to 27.8, and 30.6 to 30.8.

    Mixtures and models for ideal-gas-mixture behavior are discussed in Chapters 26 and 27, chemical reactions and chemical equilibrium in Chapters 29 and 30, and combustion in Chapter 31.

    3. Graduate Course: Thermodynamics: Foundations and Applications. This course is for seniors and first-year graduate students in the schools of science and engineering. Its purpose is to expose the student to the full generality of thermodynamics as a subject that pervades the foundations of all of physics and much of engineering. It represents an entirely new conception. It includes Chapters 2 to 13 and selected topics from Chapters 15 to 31.

    The foundations of thermodynamics are discussed in Chapters 2 to 12, and novel graphical representations on the energy versus entropy diagram in Chapter 13. Selected topics and homework problems from Chapters 15 to 31 may also serve as preparation for the doctoral qualifying exam in thermodynamics.

    4. Graduate Course: Advanced Topics in Thermodynamics. If a program includes two semesters of thermodynamics for first-year graduate students, review Chapters 14 to 25 with emphasis on the parts not covered in the first semester, and study in depth Chapters 26 to 31.

    2 Kinematics and Dynamics

    The description of physical phenomena requires rigorous consideration of many basic concepts. We assume that ideas such as space, time, reference frame, velocity, acceleration, force, particle, atom, and molecule are known from introductory courses in physics and need not be reemphasized here. On the other hand, because of the breadth and depth of the scope of our exposition, the definitions of other concepts need special emphasis.

    In this chapter we emphasize the definitions of system, property, and state, and discuss some of the ideas related to the causal laws governing motions that occur either spontaneously or because of interactions between systems. The branch of physics that describes the possible or allowed states of a system is called kinematics, and that which gives a causal description of the time evolution of a state is called dynamics.

    2.1 Systems

    In any physical study we always focus attention on a collection of constituents that are subjected to a nest of forces. When the constituents and the nest of forces are well defined, we call such a collection a system. Everything that is not included in the system we call the environment or the surroundings of the system. In this section we discuss the requirements for constituents and nest of forces to be well defined, namely, the requirements for the definition of a system.

    A constituent is either a material object, such as a molecule, atom, ion, nucleus, and elementary particle, or a field, such as the electromagnetic field, that for the given description is to be considered as an elementary building block. For each system a constituent is specified by the type and the range of values of its amount. The type may be a specific molecule, atom, ion, nucleus, elementary particle, or a specific field. For example, in certain studies of water, the type of constituent may be just the H2O molecule. In other studies of water, however, more than one type of constituent may be required, such as H2, O2, and H2O molecules, H and O atoms, H+ and O– ions, and electrons, or H3O+, OH-, and H2O molecules. Again, in studies of electromagnetic radiation, the type of constituent is the electromagnetic field.

    Example 2.1.  In each of the following studies, is the water molecule, H2O, the only type of constituent?

    (a) The study of a liquid solution of salt, NaCl, in water.

    (b) The study of absorption of photons by H 2 O molecules in a microwave oven.

    (c) The study of steam generation in a boiler.

    Solution.

    (a) No. The salt molecule is also necessary.

    (b) No. Here we must also consider the electromagnetic field.

    (c) Yes. For a wide variety of applications and range of conditions, H 2 O is the only type of molecule that need be considered.

    Example 2.2.  In each of the following studies, what are the types of constituents that must be specified?

    (a) The study of the hydrolysis of water molecules according to the reaction mechanism 2H 2 O = 2H 2 + O 2 .

    (b) The study of the oxidation of hydrogen, H 2 , in a mixture of nitrogen, N 2 , and oxygen, O 2 , according to the reaction mechanism 2H 2 + O 2 = 2H 2 O.

    (c) The study of the electrolysis of water according to the cathodic reaction mechanism 2 H 2 O = 4 H + + 4 e - + O 2 and the anodic reaction mechanism 4 H + + 4 e - =2 H 2 , where e - stands for a negatively charged electron, and H + for a positively charged hydrogen ion.

    (d) The study of pure water in the form of solid, liquid, or vapor in the absence of any chemical reaction.

    Solution.

    (a) Water molecules, H 2 O, hydrogen molecules, H 2 , and oxygen molecules, O 2 .

    (b) Same molecules as in study (a), plus nitrogen molecules, N 2 .

    (c) H 2 O, H 2 , and O 2 molecules and electrons and positively charged hydrogen atoms.

    (d) H 2 O molecules only.

    The amount of a constituent of a given type can be evaluated at any instant of time by means of counting. The value of the amount of a constituent is the numerical result of the counting procedure and is expressed as a number of particles. It may also be expressed in terms of units of mass, based on the concept of mass that we discuss in due course. For example, we may have focused attention onto a constituent consisting of water molecules, and at a given instant of time the value of the amount may be 10²⁶ molecules or about 3 kg. Again, at another instant of time the value of the amount of water may be 2 molecules or about 6 x 10-26 kg.

    The unit of measure for amounts of constituents adopted by the International System of units (SI) is the mole, denoted by the symbol mol and such that 1 mol = 6.022 x 10²³ molecules or atoms.

    When modeling a specific phenomenon, during which the amount of a constituent never takes values either below a certain lower bound, or above a certain upper bound, or both, the description is sometimes simplified if as part of the system definition we specify the range of values of the amount of that constituent, that is, we specify that the model applies only for values of the amount within the selected range. If it is not specified, we say that the range of values of an amount is unbounded, that is, the model holds for all values of the amount from zero to an indefinitely large number. For example, in certain studies of water, for each of the constituents H2, O2, and H2O, the range of values of each amount is bounded from below by a lower value, say, 20 molecules, and unbounded from above. In other studies, the ranges of values may be bounded from zero to 10 molecules for the amount of H2O, from zero to 10 molecules for the amount of H2, and from zero to five molecules for the amount of O2. Again, in some studies the lower and upper bounds coincide and we have a fixed value of the amount, that is, the range includes just a single value of the amount, such as one molecule of H2 or three atoms of H. Values of the amounts of constituents can be large or small, depending on the system under study.

    Example 2.3.  For each of the following studies, specify an appropriate range of values of the amount of H2O.

    (a) The study of the rotation of 2 kg of water in a closed, liquid-proof cylinder spinning at varying speed.

    (b) The study of the filling of an initially empty tank with water from a faucet.

    (c) The study of the electrolysis of 1 kg of water.

    (d) The study of a water molecule in a fixed-volume enclosure.

    (e) The study of water properties for all applications and ranges of conditions.

    Solution.

    (a) A fixed amount equal to 2 kg.

    (b) A range between zero and the largest amount that the tank can contain.

    (c) A range between zero and 1 kg.

    (d) A fixed amount equal to one molecule.

    (e) An unbounded range between zero and an indefinitely large value.

    The forces required to define a system are of two kinds, internal and external. Internal forces describe the specified influences that hold the molecular structure of a constituent together, such as the forces between the nuclei and the electrons of H2O, the influences between constituents, such as the forces between H2O molecules, and the forces that promote or inhibit reaction schemes by which some constituents may combine or dissociate to give rise to other constituents. They are part of the specification of the system and may differ from study to study of the same constituents. For example, in some studies of H2 and O2, the forces involved in the formation of water out of hydrogen and oxygen, the chemical reaction mechanism 2 H2 + O2 = 2 H2O, may be neglected as unimportant, whereas in other studies, these forces may be included as important. Each internal force on a given constituent depends on the coordinates of that constituent and on the coordinates of one or more of the other constituents of the system, but not on any coordinates of constituents of bodies in the environment.

    Example 2.4.  What are the internal forces for the following types of constituents?

    (a) A collection of positively charged oxygen ions, O + , and negatively charged oxygen ions, O – .

    (b) A collection of H 2 O molecules.

    Solution.

    (a) The Coulomb force on each positive and each negative charge resulting from all the other electric charges, and the forces that account for holding the parts of each of these oxygen ions together.

    (b) The forces between H 2 O molecules, and the forces that hold the protons, neutrons, and electrons of the H 2 O molecule together.

    In general, a large number of chemical and nuclear reaction mechanisms are always effective and, in principle, all should be included in the system specification. However, depending on the scope of each study, most reaction mechanisms operate on such a long time scale, that is, they are so slow that for all practical purposes it is as if they were not effective. Accordingly, we simplify the system specification by neglecting the extremely slow reaction mechanisms and concentrating our attention on the most effective ones. For example, in some studies of water we may specify that the reaction mechanism 2H2 + O2 = 2H2O is the only one to be effective. Then we deliberately neglect all the other mechanisms, such as H2 = 2H, O2 = 2O, and 2H2O = H3O+ + OH-. Of course, such a modeling assumption is reflected in and deeply affects all the results of the study.

    The concept of neglecting some internal reaction mechanisms, when they are much slower than others that dominate the phenomenon under study, has proved very practical and successful. Sometimes, this modeling concept is rendered more vividly by using terms such as passive resistances, anticatalysts, reaction inhibitors, internal constraints, or, simply, constraints, to suggest the idea that actual forces within the system prevent the slow reaction mechanisms from occurring at all and allow only the set of specified mechanisms. Of course, the concept of internal constraint is not just limited to the specification of which chemical or nuclear reaction mechanisms are relevant, but is used as well, for example, when we specify that two parts of a system are not allowed to exchange any amount of a given constituent.

    Each external force describes a well-defined influence on the constituents by bodies not included in the collection under study, such as the influence of applied gravity, electric charges, magnets, and the solid walls of the container that confines the constituents within a region of space. Each external force on a given constituent depends on the coordinates of that constituent and one or more external parameters or, simply, parameters that describe the overall effect of bodies in the environment, but not on the coordinates of any other constituent, either of the system or of bodies in the environment. For example, the effects of gravity on the water molecules in a small container depend on the elevations of these molecules above sea level, and on the intensity of gravity, but not on the coordinates of the substances of the earth. The gravitational potential γ is the parameter associated with this external gravitational force, where γ = gz, g is the constant gravitational acceleration, and z the elevation above sea level. Again, the effects of the walls of an airtight container on an enclosed gas depend on the positions of the gas molecules relative to the internal surface of the container, but not on the coordinates of the molecules of the materials of the walls. For a wide variety of applications, the effects of the walls are adequately described by the volume of the container, that is, volume is a parameter. For other applications, we may need a more detailed geometrical description of the shape of the enclosure in which the constituents are confined. For example, for a cubical enclosure of side L, the length L is a parameter of the confining external forces. More formally, the forces that confine constituents within an enclosure are described by a potential equal to zero inside the enclosed region of space, and equal to infinity outside. Thus the gradient of the potential, that is, the force field, is zero inside the enclosure and infinite at the wall so that the particles in the enclosure cannot escape.

    Each parameter, such as the gravitational potential γ, the volume V, and the length L in the three examples just cited, can be evaluated at any instant of time by means of well-defined measurement procedures. The value of a parameter is the numerical result of the corresponding measurement. For example, we may have focused attention onto a variable-volume enclosure, and at a given instant of time the value of the volume may be 2 cubic meters (m³), whereas at another instant of time it may be 3 m³. For some parameters, the description of a specific phenomenon is sometimes simplified if as part of the system definition we specify the range of values of the parameter, that is, if we restrict our attention on a range of values between a given lower bound and a given upper bound. In some studies, the lower and upper bounds of a parameter may coincide and then we have a fixed value of the parameter, such as a fixed value of the volume, V = 1 m³. Values of parameters can be large or small, depending on the system under study.

    Example 2.5.  What are the parameters for the following types of external forces?

    (a) The forces that confine constituents within a spherical enclosure of diameter D.

    (b) The forces that , confine constituents within a rectangular enclosure of sides L 1 , L 2 , and L 3 .

    (c) An applied gravity field which at elevation z has potential gz and intensity g.

    (d) An applied electrostatic field which at some location in space has potential Ψ and intensity E.

    (e) The forces that constrain constituents onto a surface characterized by a surface area a.

    Solution.

    (a) The parameter is the diameter D.

    (b) The parameters are the lengths L 1 , L 2 , and L 3 .

    (c) The parameter is the potential gz.

    (d) The parameter is the potential Ψ

    (e) The parameter is the surface area a.

    For certain systems, the characterization of the external forces may require two or more parameters. For example, in a system consisting of a fixed amount of hydrogen, half of which is confined in a volume V' and the other half in a volume V", we need the two parameters V' and V", and their respective ranges, for the characterization of the effects of the confining walls.

    Each parameter may vary either independently within its range, or be internally constrained to vary in a manner consistent with given interconnections. For example, if a system has two volumes V' and V", these volumes may be internally constrained so as to satisfy the relation V' + V" = constant.

    Each internal restriction on the changes of values of the parameters is called an internal constraint or, simply, a constraint. We use the same term as just introduced in connection with amounts of constituents and internal reaction mechanisms because also here it refers to the specification of internal restrictions that define the internal structure of the system.¹

    In summary, a system is a collection of constituents that is well defined by the following specifications: (1) the type and the range of values of the amount of each constituent, (2) the type and the range of values of each of the parameters that are needed to characterize the external forces, (3) the nature of the internal forces, including the internal reaction mechanisms, and (4) the constraints on the changes in values of amounts of constituents and parameters that the internal forces can accomplish. For a system consisting of r different types of constituents, we represent their amounts by the vector n = {nl, n2, ..., nr}, where nl denotes the amount of the first type of constituent, n2 the amount of the second type, and so on. We use the same symbols n = {nl, n2, ..., nr}, with or without an additional subscript, to denote values of the amounts of the constituents within their respective ranges. For a system with external forces that depend on s different parameters, we represent them by the vector β = {β1, β2, ..., βs}, where β1 denotes the first parameter, β2 the second, and so on. Moreover, we use the same symbols β = {β1, β2,..., βs}, with or without an additional subscript, to denote values of the parameters within their respective ranges. For some systems, the values of the amounts of constituents and the parameters, that is, the values of n1, n2, ..., nr, β1, β2, ..., βs, are all fixed. For other systems, some or all of these values may vary over bounded or unbounded ranges. Then part of the definition of the system is the specification of the lower and upper bounds to the values of each amount n1, n2, ..., nr and each parameter β1, β2, ..., βs

    Two systems are identical if they consist of the same types of constituents, experience the same internal and external forces, and have the same ranges of values of amounts of constituents and parameters, and the same constraints. If any of these identities is not valid, the two systems are different.

    An example of a well-defined system is sketched in Figure 2.1. An amount n of water molecules is confined within a watertight container of volume V. We assume that the only internal forces are the intermolecular and those that maintain the integrity of the H2O molecules, and that the external forces are those exerted by the fixed walls of the container on the H2O molecules. Another example of a system is sketched in Figure 2.2. The specifications are the same as for the system in Figure 2.1.plus a movable piston. Because the piston is movable, the volume of the container has values in the range from V1 to V2.

    A third example of a system is a thermonuclear plasma in which specified amounts of atoms and molecules experience ionizing forces and nuclear fusion reactions, release electromagnetic radiation and neutrons, and are confined by a suitably designed, time- varying electromagnetic field. Other examples of systems are the materials confined within the boundaries of the components of a pressurized-water nuclear power plant (Figure 2.3), the materials within a living cell (Figure 2.4), the neutrons within the core of a nuclear reactor, and the steam flowing through the various stages of a steam turbine. In each of the examples just cited, the amounts and types of constituents are those that exist within the boundary of the system, but the boundary may be crossed by various materials and radiation. It turns out that under certain conditions, as soon as particles enter the boundary they can be regarded as part of the system, and as soon as they exit from the boundary they are no longer part of the system but can be regarded as part of the environment.

    Figure 2.1  An amount n of water molecules confined within a watertight container of volume V.

    Figure 2.2  An amount n of water molecules confined within a watertight container of variable volume V ranging from V1 to V2.

    Figure 2.3  Schematic of components of a pressurized-water nuclear-reactor steam power plant.

    Figure 2.4  Schematic of a living cell.

    Example 2.6.  Each of two rigid bottles has the same volume V and contains 1 kg of water. In bottle A all the water is in liquid form, whereas in bottle B one half is in the form of ice and the other half in the form of vapor. If the contents of each bottle are regarded as a system, are the two systems identical or different?

    Solution.  The two systems are identical because each consists of the same type and amount of constituent and has the same internal and external forces. The form of the constituent—ice, liquid, and vapor—is not part of the system definition.

    Example 2.7.  A piece of paramagnetic material, 5 cm x 5 cm x 25 cm, is on a bench. An identical piece of the same material is placed between the poles of a very strong magnet. Can the two pieces be regarded as identical systems?

    Solution.  If the magnetic field is defined as having a specific value, the two pieces cannot be regarded as identical systems because they do not have the same external parameters: one has zero magnetic field and the other a strong magnetic field. However, if the intensity of the applied magnetic field is specified to have a range from zero to a large value at least equal to the strength of the magnet, the two pieces can be regarded as identical systems because each is composed of the same piece of matter and is subjected to forces characterized by the same parameter over the same range of values.

    Because external forces are independent of coordinates of constituents in the environment, the constituents of a system are said to be separable from the environment or, simply, separable. This distinction is not trivial. If the forces exerted by a body not included in the object of study depend explicitly on the coordinates of the constituents of that body, the object of study is not separable and cannot be a system. To proceed in this case, we must redefine the collection of constituents so as to include the body in question. Thus the troublesome forces become internal, and a system may be defined. For example, this situation arises when we wish to study the oxygen atom in a molecule of water. Such an atom cannot be defined as a system because it experiences forces that depend explicitly on the coordinates of the two hydrogen atoms in the water molecule. However, by including the two hydrogen atoms in the object of study, these forces become internal, and the oxygen and hydrogen atoms bound together in a water molecule can be well defined as a system.

    Example 2.8.  Two identical unrestrained magnets are near each other on a table but are unaffected by other magnets. Can we define one of the two magnets as a system?

    Solution. No. The magnetic force depends explicitly on the coordinates of both magnets. Hence the two magnets are not separable. To proceed, we must include both magnets in the system.

    We emphasize that, from here on, whenever we use the term system, such as when we state the laws of thermodynamics and derive all the consequent results, we imply that each system is well defined according to all the specifications and restrictions discussed in the present section. Most of the basic concepts we develop in this book, including the concepts of energy, entropy, and temperature, would be ill defined if we attempted to apply them to a collection of constituents that, for example, does not satisfy the requirement of separability just discussed. Thus, as we proceed, we must bear in mind all the restrictions that delimit the term system.

    2.2 Properties and States

    At any instant of time, the amount of each type of constituent and the parameters of each external force of a system have specific values within their respective ranges. By themselves, values for all the amounts and the parameters are not sufficient to determine all the characteristic features of the system at that time. For example, they do not fix the values of the internal and external forces because these values depend on the coordinates of the constituents. To complete the characterization of a system at an instant of time, we need not only the values of the amounts of constituents and the parameters but also the values of all the properties at that instant of time. A property is a system attribute that can be quantitatively evaluated at any instant of time by means of a set of well-defined measurements and operations performed on the system, provided that the resulting value does not depend on the measuring devices, other systems in the environment, or other instants of time.² For example, the instantaneous position of a particle within a fixed container is a property. However, the distance covered by that particle over a period of 30 minutes is not a property. As we see later, other important properties are energy, adiabatic availability, and entropy.

    The concept of velocity of classical mechanics provides a simple illustration of the measurements and operations involved in the definition of a property. We measure position x1 at the instant of interest t1, position x2 at a later time t2 = t1 + Δt, divide the vector Δx = x2 — x1 joining the two positions by Δt = t2 — t1, and find the ratio Δx/Δt. For any finite Δt, the distance Δx is finite, and does not correspond to a property because it depends on two instants of time. But if Δt is infinitesimal, that is, in the limit as t2 approaches t1 also the distance Δx becomes infinitesimal and, in general, the ratio Δx/Δt assumes a finite limiting value. This limiting value depends only on the time instant of interest, t1. It is the value of the property that we call velocity.

    Although an indefinite number of properties can be attributed to a system, some of these properties are interdependent, that is, a change in value of one such property affects the values of some other properties. For example, the speed and the square of the speed are two interdependent properties. Nevertheless, for each well-defined system it is always possible to identify many different sets of independent properties, each of which satisfies the following two conditions.

    (1) The values of properties in the set can be varied independently, that is, the value of each property in the set can be changed without affecting the value of any other property in the set.

    (2) The values of the properties in the set are sufficient to determine the values of all the other properties of the system.

    We call the properties in one such set a basis³ and denote their collection by the vector p = {p1, p2, ...}.

    For a given system, it follows that all that can be said about the results of any measurements that may be performed on the system at one instant of time is represented by: the values of the amounts of the various types of constituents, the values of the external parameters, and the values of the properties in a basis. We call the collection of all these values a possible or allowed state, or, simply, a state. Thus the state is a set that specifies everything about a system at one instant of time. The description of the states of a system is the subject of the branch of physics known as kinematics.

    Two states of a system are identical only when the values of the amounts of all the constituents, the values of all the external parameters, and the values of all the properties for one state are identical to the respective values for the other state. Otherwise, the two are different states.

    For a given system A we may use the symbol Ai to denote the ith state, (n)i to denote the set of values of all the amounts of constituents, (β)i to denote the set of values of all the external parameters, and (P)i to denote the set of values of all the properties in a basis. Thus Ai is determined by the sets (n)i, (β)i, (P)i. Two states A1 and A2 are identical if each entry in (n)1, (β)1, (P)1 is equal to the corresponding entry in (n)2, (β)2, (P)2.

    Example 2.9.  A system consists of water molecules contained in an injection syringe with a volume range from 0 to 5 cubic centimeters (cm³). Do any of the following four descriptions specify the same state? The amount of water in the syringe and its volume are, respectively: (a) 2 g and 3 cm³, (b) 2 g and 2 cm³, (c) 3 g and 3 cm³, and (d) 2 g and 3 cm³.

    Solution.  Descriptions (a) and (b) specify different states because the value of the parameter volume in (a) is different from that in (b). Descriptions (a) and (c) specify different states because the value of the amount of water in (a) is different from that in (c). Descriptions (b) and (c) specify different states because the values of the amount of water, and the parameter volume in (b) are different from those in (c). Descriptions (a) and (d) specify that the two states have identical values of amount of water, and identical values of the parameter volume, but we cannot conclude that they specify the same state because we are not told whether the values of all properties of (a) are identical

    Enjoying the preview?
    Page 1 of 1