Vous êtes sur la page 1sur 65
Supplementary Chapter on Thermodynamics for J.F. Shackelford, Introduction to Materials Science for Engineers, 6th Edition, Prentice- Hall, Upper Saddle River, NJ, 2005 The following chapter on the subject of “thermodynamics” appeared in the Instructor’s Manual for the 5th Edition of Introduction to Materials Science for Engineers. The chapter is reproduced here in its entirety. All cross-references to the Sth Edition are also valid for the 6th Edition. Table of Contents 1 The First Law 2 The Second Law 3. The Third Law 4 The Free Energies 5 Free Energy and Phase Equilibria 6 The Thermodynamics of Surfaces 7 The Thermodynamics of Kinetics 8 — The Thermodynamics of Environmental Reactions 510 516 527 529 533 552 559 564 As the name implies, thermodynamics focuses on the nature of heat as a form of energy and the limitations with which heat can be converted to other forms of energy. ‘These principles have wide-ranging implications for the practical applications of engineering materials. ‘The foundation of thermodynamics is a set of three physical laws based on various experimental observations of the behavior of materials. The first law is a formal statement of the conservation of energy. An important aspect of this law is that internal energy is a “state” function, i.e., the magnitude of its change in a process is dependent only on the beginning and ending states and not on the path joining those states. In addition to internal energy, two other thermodynamic variables will be defined relative to the first law discussion. Enthalpy is the energy of a system plus the product of its pressure and volume. Heat capacity is the derivative of heat input to a material with respect to its temperature rise. The second law deals with the practical limitations on the conversion of heat to work. Tt leads to the definition of yet another thermodynamic state function, entropy. A compact statement of the second law is that any spontaneous process leads to an increase in entropy. The second law can also be stated in terms of the statistical nature of the atomic-scale structure of materials. the third law states that the entropy of a pure material goes to zero when its temperature has been reduced to absolute zero. Free energy is @ thermodynamic state function that is minimized at equilibrium. as such, the free energy provides the theoretical foundation for the nature of phase diagrams (see Chapter 9 in Shackelford, 5th Edition). Thermodynamics can provide a fuller understanding of the significance of surfaces in naterials processes. Also, some of the thermodynamic state functions defined in this chapter (enthalpy, entropy, and free energy) are intimately associated with the kinetics of rate processes (see Chapters 5 and 10 in Shackelford, 5th Edition). Finally, thermodynamics principles lay a foundation for better understanding of the nature of environmental reactions (see Chapter 19 in Shackelford, 5th Edition). 509 1 THE FIRST LAW If a given system takes in a quantity of heat, Q, while doing an amount of work, W, there will be a net change in internal energy, AU, for that system given by au = Qg- (a1) This simple statement of the conservation of energy” is the first law of ‘thermodynamics. Although simple, it is a powerful and useful concept. one should especially note the sign convention of Equation 1-1. A heat flow to the system is positive (and from the system is negative). Conversely, work done by the system is positive (and on the system is negative). This is the so-called “engineering convention," in that the sign for work is determined by the practical perspective of the user who wishes to extract work from the system. The reader may encounter other books dealing with thermodynamics (especially in the field of chemistry) in which the sign for work is deter- mined by the perspective of the system, rather than the external user of energy. (In those books, the first law would be written AU = Q + W). Relative to Equation 1-1, it is important to note that the internal energy, U, is a state variable, i.e., a property whose value depends solely on the state of the system (established by pressure, temperature, etc.) and not by the path in reaching that state. On the other hand, heat, Q, and work, W, are not state variables. ‘Their values are functions of the path of the process in going between an initial and a final state. The value of SU, conversely, is not a function of the path, and depends only on the values of U at the initial and final state. Equation 1-1 describes energy terms associated with an overall process going between an initial and a final state. In many cases, we shall want to monitor the specific progress along a process path. For this purpose, we require a differential form of the first law involving the relationship between the incremental change in internal energy, aU, and the incremental *For our purposes in this book, we shall not have to deal with the additional consideration of the interconversion of energy and mass (E = mc”), which occurs to a significant degree in certain relativistic processes. 510 values of heat, 69, and work, 6: au = 69 - ow. (1-2) One should note that d is used to designate a differential quantity of a state variable, and 6 is used to designate a differential quantity of a path-dependant variable. In applying the first law of thermodynamics, it is convenient to define @ quantity known as the heat capacity to be the ratio of the incremental heat addition to (or loss from) a system to the corresponding incremental ‘temperature change, af: = @ e+. (1-3) Note that temperature is a state variable, This is also true for sone other comson temas such as pressure, p, and volume, V. the heat capacity 1a « most useful term, but, in general, we need to specify the conditions under which it is measured. When the system is constrained to a constant volume during the course of the measurement, the heat capacity at constant volume is designated as 5Q. a «= (2) y (4) ven the system is beld at a constant pressure, the heat capacity 4s 82 -! (Be ee An important feature of the heat capacity is that it can be related to first law terms. Consider, for example, the common case in which the work term in the first law is strictly mechanical (i.e., such terms as magnetic work are not involved). Mechanical work is produced by a force, F, acting through a distance, x. If the process is carried out relatively slowly with an absence of frictional losses, the incremental work term is ow = Fax. (1-6) Such an ideal process is called reversible. It represents the maximum efficiency for such mechanical action, as well as the maintenance of equi- librium at each state of the overall process. The term "reversible" is used because an infinitesimal force applied in an opposing direction can reverse the system from expansion to compression, and vice versa. The force is equal su to the system pressure, p, times the cross-sectional area, A, of the system (normal to the x-direction): Fax = padx. (a-7) In turn, the product of area and incremental distance represents an incre- mental volume: pAdx = av, (1-8) allowing us to rewrite Equation 1-2 as au = 69 - pav (1-9) or, 6Q = aU + pav. (1-10) A consequence of this relationship is that the heat capacity at constant volume (Equation 1-4) can be expressed in terms of internal energy, i.e., =(8) . a) 7 « (B). (Ble au Similarly, the heat capacity at constant pressure (Equation 1-5) can be rewritten as: 3 2), - C=) "5 (2-12) P ‘The combination of internal energy. and a pressure-volume product occurs frequently in thermodynamic analyses. As a result, it is convenient to replace this combination of state variables with a single term, H, which is, in turn, a state variable: # v+py (2-13) and is called the enthalpy (from the Greek word thalpein, meaning “to heat"). Although the enthalpy may strike the beginning student as a somewhat abstract concept, the direct connection to a practical engineering measure- ment (of heat capacity) can be seen by combining Equations 1-12 and 1-13: 2) c. =(#),. a4) P e P ae In the latter part of this chapter, we shall find numerous other examples of the practical utility of enthalpy as a thermodynamic variable. 512 SAMPLE PROBLEM 1-1 The heat capacity of a material is a function of temperature. For typical solids, the heat capacity at constant pressure, Cy, can be expressed by an empirical relation of the form: Cc, = atbre or > T/(mol+K). For silver over the temperature range of 298 K to its melting point (2235 &), a = 21.1 J/(mol+k) b = 8.55 x 10°? 3/(mol+K?) © = 1.50 x 10° (g-K)/mol. Calculate the heat input necessary to reversibly heat a kg of silver from 25°C to 500°C under a constant pressure of 1 atm. SOLUTION Rearranging Equation 1-5, = ca (20), = 6a a, 2 a= f? cen. 1 Using the given empirical relation for C,, 2) ar 2 a= [2 (as prs of” y 513 For this specific case, note that 1, = (25 + 273)K = 298K and Tz = (500 + 273)K = 773 K, and that the C, parameters are given above. Then, Q = 21.2 J/(mol+K) (773 K - 298 R) + 25x30" aftworee?) (73 1)? - (298)? + 1.50 x 20%(exy/nor ((773 xy°? - (298 7) 3 = 12.6 x 10° J/mol. Finally, we can use Appendix 1 of Shackelford, sth zdition, to obtain the result for 1 kg of silve: . 3 1 mol 1000 _¢ Q = 12.6% 10° g/mol x TTS x = 116 x 107 a/kg. 514 SAMPLE PROBLEM 1-2 What is the enthalpy of 1 kg of silver at 500°C relative to its value at 25°C? ‘SOLUTION Rearranging Equation 1-14, = (38) car or 2 1, Joa - [Poa > 1 2 giving 2, 2 H(t) HCN) = I, cat. For the case at hand, 113K (S00%c) - w (25ec) = f cen. 298 K The integral on the right side of the above equation was deter- mined in Sample Problem 1-1. ‘Therefore, H (500°C) - H (258C) = 116 x 10° a/kg. 515 2 ‘THE SECOND LAW The first law is useful as a means for doing the necessary "book- keeping" on the energy terms associated with a given process. ‘The second law provides a statement as to the limitations on the conversion of heat to work. A convenient means for introducing a formal statement of the second law is to describe the nature of a simple heat engine. Figure 2-1 shows a schematic illustration of a Carnot* cycle involving the expansion and compression of an ideal gas within a piston. ‘There are four steps in the overall cycle: I. A reversible, isothermal heat addition. II. A reversible, adiabatic (i.e., perfectly insulated from any heat flow) expansion. III. A reversible, isothermal heat rejection. IV. A reversible, adiabatic compression of the gas back to the. initial state of the system (at a pressure, temperature, and volume corre- sponding to the start of step I). Figure 2-2 is a plot of the Carnot cycle on a pressure-volume @iagram. For the overall cycle (starting at point a and returning to point a), the change in internal energy (or any other state variable) mist equal zero, dee, Aus: 0. (2-1) The first law (Equation 1-1) for the overall cycle is, then, AU = 0 = + Qy - Meotale (2-2) or ¥, feotaleg or rar a where Weoce: is the net useful work extracted from the cycle and Q. and *Nicholas Leonard Sadi Carnot (1796-1832), French physicist. ‘The son of a prominent scientist and politician, Sadi Carnot followed along similar paths, including a successful military career. In analyzing engines, he introduced the concepts of a “complete cycle” and reversibility. Scarlet fever led to his untimely death at the age of 36. 516 Step | (with Q, added ) FIGURE 2-1 Schematic of a Carnot cycle in which four expansion and compres- sion steps bring the gas inside a piston-cylinder back to its original state. FIGURE 2-2 The Carnot cycle as plotted on a pressure (p) - volume (¥) diagram. 517 Qees are the heat addition (during step I) and heat rejection (during step III). We can now define an overall cycle efficiency, n, as the net work extracted divided by the total thermal energy input: Meotat | % * Sr ne % (2-4) The denominator of Equation 2-4 contains only the heat addition term, Q:. In this simple engine, the rejected heat, Qrz, is not reclaimed and does not reduce the magnitude of total thermal energy input. Because Qs2z is a negative quantity and appears in the numerator of Equation 2-4, the offi- ciency, n, of the Carnot cycle is less than unity. ‘the fact that the thermal energy, Qx, cannot be completely converted to usable work is closely related to the second law of thermodynamics. 1o quantify the cycle effi- ciency, it is useful to introduce a new state variable which will be central in a formal statement of the second law. It can be demonstrated experimentally that, for cycles such as that illustrated in Figures 2-1 and 2-2, PRs (2-5) Where the integral is taken over the complete cycle. The equality in Equation 2-5 corresponds to a completely reversible cycle. For this limiting case, the differential of reversible heat flow divided by temperature (8Q-4-/T) behaves as the differential of a state variable. This state variable is called entropy, S (from the Greek word trepein, meaning "to turn or change d, dees, as : (2-6) Returning to the Carnot cycle, we can determine an entropy change for each step: --, (2-7) As. ar Or (2-8) 518 ee f Brey Sir III ey and asy pe =o. (2-10) a Because the overall entropy change for the reversible cycle is zero, we can write that 48,4. 7 0 = As, +4 aoa + OS. + AS (2-11) Sty ‘III Tv, Combining Equation 2-11 with Equations 2-7 to 2-10 gives 2, bo pH so (2-12) The Carnot cycle efficiency (Equation 2-4) can now be rewritten: Q + (O62, 'rrr/T) (2-13) : (2-14) This result is both simple and convenient, in that it permits the cycle efficiency to be calculated with knowledge only of the operating temperature range of the engine. Also, one should note that Equation 2-14 has signifi- cance well beyond a by-product of our introduction of the second law of thermodynamics. It is a convenient approximation to the efficiency for many Practical power generation systems. As most engines will have a lower operating temperature (T,zr) near room temperature, efficiency will generally be limited by the magnitude of the upper operating temperature (Tz). This is the central reason for the extensive development of new ceramics that can withstand higher operating temperatures’ than metals in engine designs (see Chapter 12 in Shackelford, Sth Edition). In effect, Equation 2-14 tells us that heat cannot be converted to work with 100 percent efficiency. ‘This is equivalent to one of the early statements of the second law of thermodynamics, It is also closely con- 519 nected to the common observation that heat will not be spontaneously trans- ferred from a cooler to a hotter body, another early statement of the second law. A consequence of Equation 2-5 is that, for a given incremental process between two adjacent states, the heat flow for a reversible step will be greater than for an irreversible step: Boy > 62. rev (2-15) ‘irrev" Also, this means that the actual heat transfer for any incremental process can never be greater than the reversible heat transfer between the adjacent states, i.e., 8 < OO. (2-16) However, the entropy change for the system can always be calculated from Equation 2-6 using the heat transfer for the reversible path, i. 69, rev =, (2-17) The entropy change for the system's surroundings can be calculated as the reversible absorption of the actual heat lost from the system (-69), i. = : Sour Le (2-18) The net entropy change for the universe (i.e. ‘then: system + surroundings) is, as, = 48,0 + as - Bro (2-19) net. sys * Ssurr where the relative magnitudes of 6Q,., and 6Q are given by Equation 2-16. Equation 2-19 implies that any spontaneous process will be accompanied by an increase in total entropy and that entropy is maximized at equilibrium. ‘These two implications’ are our most general statements of the second law and illustrate the central role of entropy (comparable to the roles of energy and enthalpy in the first law statements). The utility of entropy as a state variable is illustrated in Figure 2-3, in which the Carnot cycle is plotted on a temperature-entropy (-S) diagram. In contrast to the p-V @iagram of Figure 2-2, the 7-8 520 s FIGURE 2-3 The Carnot cycle as plotted on a temperature (7) - e1 itropy (5) diagram. Note that the area bounded by the plot a-b-c-d is equal to the work extracted from the engine in one cycle. (This is also true for the p-V plot in Figure 2-2.) diagram has a simple rectangular geometry. Another convenient feature of the T-S plot is that the area enclosed by the plot a-b-c-d is equal to the work extracted from the engine (Weoca:) in one cycle. A careful inspection of Equations 2-3, 2-7, 2-8, 2-9, and 2-10. confirms this fact. The area enclosed by the p-V plot in Figure 2-2 is also equal to Worai, although not as simple to evaluate as the rectangular area in Figure 2-3. our discussion of thermodynamics to this point has dealt strictly with macroscopic observations. Such macroscopic properties are based on micro- scopic (or submicroscopic) structure (see Section 1.4 in Shackelford, Sth Edition). The nature of the second law statements following from Rqua- tion 2-19 can be related to the statistical probability of the atomic arrangement of the system in question. Specifically, the entropy of the system is related to the number of possible atomic arrangements, W, by s = kinwW (2-20) where k is Boltzmann's constant. The development of expressions such as Equation 2-20 is appropriately termed "statistical thermodynamics." The lassical thermodynamics." Most of our coverage in this Manuel | will be limited, for the sake of time and macroscopic treatment is termed " sai space, to classical thermodynamics but, even in an introductory materials science book, there will be occasions in which the statistical treatment is advantageous. The entropy of mixing for solid solutions is an example. First, we shall take Equation 2-20 as a valid statement of the value of entropy without reviewing the experimental evidence that confirms it. More detailed references, such as those at the ond of PartII, are available for such purposes. Of more immediate importance, Equation 2-20 provides an additional statement of the second law, viz., a spontaneous process is one in which the system goes to a more probable state (larger value of W). Consider, now, the specific case of a substitutional solid solution (note Figure 4-2 in Shackelford, Sth Edition). If there are N, total atomic sites and n atoms of element A, there will be (No-n) atoms of element B in a binary system (such as the copper-nickel solid solution). The number of distinct ways in which the A and B atoms can be arranged on the No sites is (again, without derivation): N WS GPA (2-21) Therefore, the entropy of mixing, AS,, for the solid solution is Not os, = loa | (2-22) Because the numbers of atoms are large (on the order of Avogadro's number, in typical cases), we can replace the factorial terms using Stirling's® approximation: InN! = NinN-W (2-23) where N is any “large” number. Applying Stirling's approximation to Equa~ tion 2-22 gives as, = (2-24) SJames Stirling (1692-1770), Scottish mathematician. Stirling's significant contributions to mathematics were confined to the 1720s and 1730s. His ability for practical financial calculations led to his being asked to reorganize the Scottish Mining Company in the lead mines of Lead- hills, Lanarkshire. As a successful administrator, he spent most of his time after 1735 in that renote village. 522 Noting that the atom fraction of element A, X,, is n/No and the atom fraction of element B, Xm, is (No-n)/No, Equation 2-24 can be rewritten as Os, = -Nok (%, In X, +X, In x]. (2-25) For 1 mole of solid solution (No = Navegaace): as, = -R [X, Ink, +X, In X) (2-26) where R is the universal gas constant (= Navk). Figure 2-4 is a plot of AS, over the composition range for a binary solid solution system. A general feature of this plot is that the entropy increases with the increasing degree of "disorder" represented by the random mixing of unlike atom: XB FIGURE 2-4 The entropy of mixing (of atoms A and B in a substitutional solid solution) as a function of composition (atomic fraction of B). This is a plot of Equation 2-26. 523 SAMPLE PROBLEM 2-1 (a) Calculate the maximum (Carnot) efficiency for a conventional automobile engine (made largely from iron and steel) operating at a maximum temperature of 800°C. (b) Repeat (a) for a prototype Si,N, ceramic engine operating at @ maximum temperature of 1200°C. SOLUTION (a) We can use Equation 2-14 with the assumption that the lower operating temperature is room temperature (taken as 25°C = 298 K). For an upper operating temperature of 600°C.= 1073 K, Ty ~ Tir er 1073 - 298 OTS = «0-72: 72% (b) For an upper operating temperature of 1200°C = 1473 K, 1473 = 298 Sas 0.80 or 80% 524 SAMPLE PROBLEM 2-2 Calculate the percent error in using Stirling's approximation (Equation 2-23) for (a) N= 5, (b) N= 10, and (c) N= 50, SOLUTION (a) For N= 5, N!= 120 and InN! = 4.787. But, NinN-N = SinS-S = 3.047. Therefore, Percent error 36.3%. (b) An 101 = 1n 3,628,800 = 15.104 10 In 10 - 10 = 13.026 = 15.104 ~ 13.026 : Percent error = 28:104—1 x 1008 = 13.88. (c) In 50! = In 3.04 x 10 = 148.5 50 In 50 - 50 = 145.6 68.5 - 145.6 . Percent error = $48:5-=—145:6, 094 = 1,958, 525 SAMPLE PROBLEM 2-3 Calculate the entropy of mixing for 100 g of an equimolar solid solution of Cu and Ni. SOLUTION First, we must determine the number of moles of solution. Taking Xx g of Cu and (100 - x) g of Ni, and using data from Appendix 1 of Shackelford, 5th Edition, os Et = no. moles Cu = 23735 g/mol ~ @3.55 Tt ana no. = GW-2g - moles Mi = 55.77 g/mol 2 As the alloy is equimolar, 100 - 58.71 or x = 52.09 ana 100-x = 48.09. The total number of moles will be 7 = 52.0 48.0. | Peot ~ Tout "ui “ 63.55 * 58.77 7 1-64 mol. Turning to Equation 2-26, we can calculate AS, As, = (1.64 mol)(-8.314 J/(mol+K)(0.5 1n 0.5 + 0.5 1n 0.5) 9.45 J/K. 526 3 THE THIRD LAW The Carnot cycle of Figures 2-1, 2-2, and 2-3 provides for a definition of an absolute temperature scale. For the overall cycle, we can recall that 2 Q =, Sm. 4 (2-12) fen or, by rearranging e-o. (3-1) This relationship establishes the temperature scale by the ratio of heat input to outflow. By assigning a value to some convenient reference point (e-g., 273.16 K for the freezing point of pure water at 1 atm pressure), all other temperatures follow from Equation 3-1. Returning to Equation 2-6, we can integrate that expression to obtain values of entropy at a given temperature, T', and pressure, p (allowing us to use Equation 1-5): tT! = fe : en I, Patt sn. (3-2) The value of entropy at absolute zero is a constant of integration. In the early part of this century, it was demonstrated that the entropy of all pure materials approaches the same value at 0 K. or the convenience of being able to calculate absolute values of entropy at various temperatures, the value of entropy (for any material) at absolute zero is arbitrarily taken as zero, i.e, Spep = (3-3) This definition is the third law of thermodynamics. Equation 3-2 can, then, be simplified to allow for the calculation of the absolute entropy of a system at temperature T (3-4) 527 SAMPLE PROBLEM 3-1 Calculate the entropy change that occurs upon heating 1 kg of silver from 25°C to 500°C under a constant pressure of 1 atm. SOLUTION This can be calculated with the use of a modified form of Equa- tion 3-4, viz., oorc 85 = Ss.000 - Base = | = ar. soore ~ 82590 = Jace The heat capacity data for this case was given in Sample Problem 1-1, i.e., orc a as = Ke tbter?)ar 25°c a sooec = 773 K = alnt+pr-€ asec = 298 K . ax), 7 = aln (BB) ) >’ (Ga K) = (298 ») $ (73 w) - (290 0) 798 K 5, x (773 8)-(296 x) ~ (Es80.x 20%(a-K) no 21.1 J/(mol+K) In (HR) (2.58 x 10 3/(ao1+x2)) (0 x) * ~ (298 07) 25.0 3/(mol+k). As in Sample Problem 1-1, we finally need to convert to a basis of 1 kg: 1 mol, 1000 ¢ AS = 39.5 S/(moleK) x Gp7gg x “YE 232 (3/K)/kg. 528 4 ‘THE FREE ENERGIES In stating the three laws of thermodynamics, the variables internal energy, enthalpy, and entropy have been related to processes that may or may not have been spontaneous. Our next goal is to define an additional variable that can predict whether or not a given process is spontaneous. ‘This is comparable to having, in the field of mechanics, a potential energy function which is minimized at equilibrium. A clue as to the form of a “thermodynamic potential" variable comes from reflection on the Carnot cycle introduced in Section 2. The heat rejected in Step III of Figure 2-1 was associated with the limit to engine efficiency (representing, in effect, “unavailable work"). As Qecr = Tez: 4Szzz (by rearranging Equation 2-9), we can propose a new variable, F, which is the energy of the system, U, reduced by a TS produc Fos uU-ts. (4-1) The differential of F will be ar = dU - tds - sar. (4-2) Using the first law (Equation 1-2), we can modify Equation 4-2 to become ar = 69 - 6W- Tds - sar. (4-3) For a process occurring at constant volume and temperature, 6W=0 (because av = 0) and, of course, a? = 0, giving the result: (yy = 8Q- Tes. (4-4) The differential entropy term can be replaced by the equality in Equa- tion 2-6, giving (yg = 82 - Baye (4-5) The relationship in Equation 2-16 (89 $ 69...) indicates that Equation 4-5 can be restated as: (®yy s 0 (4-6) where the equality applies to a reversible process and the inequality to an irreversible, or spontaneous, process. Equation 4-6 serves precisely the Purpose stated as our goal, viz., a variable which is minimized at 529 equilibrium. As the original definition of the variable (Equation 4-1) involved the internal energy minus an “unavailable work" term (TS), F is called the Helmholtz* free energy, which is, as shown by Equation 4-6, minimized at equilibrium for processes occurring at constant volume an@ temperature. For materials applications, processes occurring at constant Pressure and temperature are more common. For such cases, the Gibbs" free energy, G, is the appropriate variable: G = H- Ts. (4-7) Recalling the definition of enthalpy (Equation 1-13), we see that the Gibbs free energy is: G = U+py-as, (4-8) The differential of G is 4G = aU + pav - vép - Tas - sar. (4-9) Introducing Equation 1-2 gives 4G = 69 - 6H + pav ~ vap - Tas - sar. (4-10) Considering only mechanical work (6W = paV) simplifies the expression for ac, viz., 4G = 60 + Vép - tas - sar. (4-12) Introducing Equation 2-6 gives aG = 69+ Vép - 89... - sat. (4-12) “Herman Von Helmholtz (1821-1894), German physiologist and physicist. Finances kept Helmholtz from studying physics as a young man. He obtained a medical degree because of state financial aid and helped to develop the "mechanistic" school of physiology. (Note Adolf Fick's related biography in Chapter. 5 of Shackelford, Sth udition.) Excellent research on physiologi- cal acoustics and optics led to his being given the prestigious chair of physics at the University of Berlin in 1866. Heinrich Hertz (whose biography appears in Chapter 18 of Shackelford, Sth Edition) was one of his students at Berlin. : “Josiah Willard Gibbs (1839-1903), American physicist. His biography appears in Chapter 9 of Shackelford, Sth Edition. 530 At constant pressure and temperature, Equation 4-12 reduces to (28), = 82 - O64 (4-13) ‘The conclusion comparable to Equation 4-6 is, then, so. (4-14) For the remainder of Part I, we will frequently use the Gibbs free energy to analyze problems of practical importance in materials engineering. 531 ‘SAMPLE PROBLEM 4-1 (a) In Sample Problem 2-3, the entropy of mixing was calculated for 100 g of an equimolar solid solution of Cu and Ni. What would be the Gibbs free energy of mixing for this alloy at 100°C, assuming an ideal solution? (We shall see in the next section that H = 0 for an ideal solution.) (b) Repeat (a) at so0°c. ‘SOLUTION (a) Equation 4-7 indicates that for a constant temperature process, AG = AH- TAS, For an ideal solution, AH = 0, giving aG = -TAs. Using the result of Sample Problem 2-3, A (1000 + 273) K (9.45 3/K) ‘S1000%¢ 712.0 ka. (b) at Sooec. AGgqgog = ~(500 + 273) K (9.45 3/K) +30 ks. Note: The negative values for these answers indicate that the mixing is spontaneous. 532 S FREE ENERGY AND PHASE EQUILIBRIA In 1876, J. Willard Gibbs developed a thermodynamic statement of fundamental importance to the practical understanding of phase diagrams. To illustrate his development, consider a simple case of two phases, a and 8, each containing different amounts of two components, 1 and 2. Next, consider the transfer of a small amount of componant 1 from phase a to phase 8. ‘The effect on the Gibbs free energy of phase a will be: (5-1) where dG, is the change in free energy of a, and aN,~ is the small amount of component 1 being removed froma. From the perspective of phase 8, aa, a (5-2) But, aN,° = -GN,*. Also, the total free energy change for the total system (a +8) must be zero due to the condition of thermodynamic equilibrium. Therefore, eotar ~ a, + a, a6, : - =o (5-3) or ag. ag, 3 -—4. (5-4) @ 3 an,’ an, In general, if we have C components and P phases, a set of equilibrium equations of the form of Equation 5-4 will apply: 533 Snr ag TE an an, - 2S : IIt a ee = an ag 3Gyy 8 Gray " = ont = yale see eee = (5-5) a 3 3 3 - 28a) PSL _ fii Tir cts a8, an, an, Each of the C rows of equations has P-1 equations for a total of C(P-1) equations associated with the free energy criterion. Tt is also true that the sum of mole fractions, X, for any given phase must be equal to one, i.e., I I axl Bt ee eK 1 I, yi We % ee 1 IT, IIT, UII. . + seen ted (5-6) Py yg? ee x Re ee 1 There are, then, P equations associated with the composition criterion. The number of degrees of freedom (independent variables) for this thermodynamic system will be the total number of variables less the nunber of equations relating them. The thermodynamic variables in the present case 534 are temperature, pressure, and composition of each phase. Temperature and pressure together are, of course, two variables. ‘There are CP composition variables (C variables for each of P phases). The total number of variables is, then, CP +2. The degrees of freedom, F, will then be CP +2 less the number of equations (C(P-1)) in Equation 5-5 and less the number of equa- tions (P) in Equation 5-6: Fo= c@P+2-c(P-1)-2 (s-7) “ Fo= c-P+2. (5-8) This is the Gibbs phase rule given without derivation as equation 9.1 in Shackelford, Sth Edition. the application of the phase rule to phase diagrams is discussed there. Now, we shall proceed with a discussion of the graphical relationship between free energy and phase diagrams A powerful feature of the free energy is its ability to indicate the equilibrium state for a material. Figure 5-1 illustrates this for a system common to our experience, H.0 at 1 atm pressure. Below 0°C, solid ice is in thermodynamic equilibrium, and the Gibbs free energy of this solid phase is lower than that for liquid or gas phases. (We shall define a phase as a chemically homogeneous portion of a system. For example, we shall commonly encounter solid phases that are chemically homogeneous but not structurally homogeneous, e.g., polycrystalline solids.) In that temperature region, the liguid and gas phases can be said to be "thermodynamically unstable" relative to the solid phase. At precisely 0°C (the melting point for ice at 1 atm pressure), the Gibbs free energies of the solid and liquid phases are equal and these two phases are in equilibrium, i.e., both are stable. In the region 0°C < T < 100°C, the Gibbs free energy for the liquid phase is lowest and that phase is stable. At 100°C (the boiling point for water at 1 atm), the liquid and gas phases are in equilibrium. Above 100°C, the gas phase is thermodynamically stable. For pure iron at 1 atm pressure, the plot of Gibbs free energy versus temperature (Figure 5-2) is somewhat more complex than that for H.0. The free energy curves for the body-centered cubic (bec) and face-centered cubic (fcc) phases of iron are close to each other, with the bec curve having somewhat greater curvature. The result is that, below 910°C, the bec structure is thermodynamically stable. Heating to 910°C produces a phase 535 0 100 ae) FIGURE 5-1 Schematic plot of the Gibbs free energy, G, as a function of temperature, T, for various phases of H,0 at 1 atm pressure. The thermodynamically stable phase at a given temperature is the one with the lowest free energy. 910 1394 1538 FIGURE 5-2 Schematic plot of the Gibbs free energy, G, as a function of temperature, T, for the various phases of pure iron at 1 atm pressure. 536 transformation, as the Gibbs free energy of the fcc phase becomes equal to that of the bcc phase. Between 910°C and 1394°C, the fcc phase is thermo- dynamically stable. Between 1394°C and 1538°C, the bec phase again is stable. At 1538°C, the iron melts (as the Gibbs free energies of the bec phase and the liquid phase are equal) and, of course, the liquid phase is stable above 1538°C. In Chapter 9 of Shackelford, Sth zdition, the practical result of the free energy plots of Figures 5-1 and 5-2 is seen to be one-component phase diagrams ("maps" summarizing the experimental observa- tion of microstructures associated with certain state variables such as temperature). The phase diagram for HjO is shown in Figure 9-3 and that for iron in Figure 9-4 of Shackelford, 5th Edition. Engineering materials generally contain impurities for practical purposes. In order to understand the phase diagrams for such systems, we must consider the effect of composition on the Gibbs free energy. the basic principles can be illustrated by looking at binary systems, i.e., materials composed of two chemical components. These components can be elements (e.g., in the case of metal alloys) or chemical compounds (e.g., in the case of ceramics). We have already laid the groundwork for this discussion by developing an expression for the entropy of mixing of two elements in a metal alloy (Equation 2-26 and Figure 2-4). In Figure 5-3, we consider this simple binary alloy further. In Figure 5-3(a), we note that the absolute entropy of each component (S, and Sq) will, in general, be different. Without atomic- scale mixing, the entropy of some combination of components A and B would simply be a weighted average, as represented by the straight, dashed line in Figure 5-3(a). The entropy of mixing term (Figure 5-3(b)) must be added, giving the plot of the total entropy as a function of alloy composition in Figure 5-3(c). ‘The substantial curvature of the entropy plot (Figure 5-3(c)) is present even for an ideal solution of A and B atoms. Such ideality can be defined as one in which the average chemical bond strength between like atoms (A-A or B-B) is equivalent to that between unlike atoms (A-B). As enthalpy is directly related to energy, a plot of enthalpy versus composition for an ideal solution will be a straight line (Figure 5-4). stated in another way, the enthalpy of mixing, 4x,, for an ideal solution is zero over the composi- tion range. The resulting plot of Gibbs free energy (G =H - 7S) with composition is shown in Figure 5-5. For this ideal solution case, the components A and B can dissolve in each other in all proportions. The only 537 A B FIGURE 5-3 (a) Total entropy for a binary system without mixing. (b) The entropy of mixing (see Figure 2-4). (c) the total entropy with mixing. 538 Ha 0 1 XB | A B FIGURE 5-4 The enthalpy of an ideal solid solution. question we need to answer about the phase equilibria is whether the system exists as a solid solution or a liquid solution. the free energy helps to provide an answer: either or both. fo illustrate this, consider a plot of the Gibbs free energy for the solid solution and for the liquid solution at some temperature below the melting. point of both components A and B (Pigure 5-6). the free energy of the solid is lower at all compositions and, so, the solid solution exists for any composition. Conversely, consider the case for a temperature above the melting point of both components (Pigure 5-7). At this temperature, the situation is reversed, and the Gibbs free energy for the stable liguid solution is lower for all composi- tions. At a temperature intermediate between the melting point of the two components A and B, the liquid and solid free energy curves overlap (Figure 5-8). As the stable system will be one with the lowest free enersy, 539 G, H, TS GsH-TS x A c B FIGURE 5-5 Combining a plot of enthalpy, H, for an ideal solution (see Figure 5-4) with a plot of a -TS product (note Figure 5-3) gives a plot of Gibbs free energy (G =H - TS) with composi- tion. 540, T1Tm,A.Tm,B G XB o FIGURE 5-7 At a temperature (T,) above the melting points of bot and th pure A pare B, the Gibbs free energy of the liquid is below that for the solid at all compositions. 542 the liquid phase will be stable from a composition of X= 0 (pure A) up to point %_=X,. Similarly, the solid phase will be stable from a composition of Xa = Kz up to %_= 1 (pure B). However, between X: < Xn < Xe, the free energy of either phase will not be as low as that of a two-phase mixture of a Liquid (of composition X,) and a solid (of composition X.) represented by the Gashed line tangent in Figure 5-8. The nature of Figures 5-6, 5-7, and 5-8 leads to a binary phase diagram with complete solid solution (Figure 5-9). (the practical analysis of this diagram is treated in Section 9.2 of Shackelford, Sth gdition). Solid solutions are frequently non-ideal. An important example is the case in which the bonding strength of like atoms (A-A or B-B pairs) is, on the average, greater than for unlike atoms (A-B) pairs. In this situation, there is a positive enthalpy of mixing, AH,. This is shown in the enthalpy versus composition plot of Figure 5-10 and can be contrasted with the straight-line plot of Figure 5-4. The practical result of this enthalpy behavior is that, at relatively low temperatures, separate A-rich and B-rich solid solutions tend to exist rather than a single, homogeneous solid solution at all compositions. why this is so can be appreciated by combin- ing the enthalpy plot of Figure 5-10 with an entropy plot to give a net plot of Gibbs free energy with composition (Figure 5-11(a)). Concentrating on the free energy plot in Figure 5-11(b), we again encounter a case (similar to Figure S-8) where a dashed-line. tangent can provide a lower free energy for the system than the solid line plot. In this case, a two-phase mixture of solid solutions with compositions X, and X, will be stable for an overall system composition in the range X, FIGURE 7-2 The activation enthalpy, AH, for the atom motion described in Figure 7-1 is equivalent to the “activation energy" of Figure 5-2 in Shackelford, Sth zéition. 561 SAMPLE PROBLEM 7-1 A theoretical model of the interstitial diffusion of carbon in a bee metal gives an expression: Dew Belvo sore where a is the lattice parameter and v the lattice vibrational frequen- cy. Using the experimental data of Table 5.2 in Shackelford, Sth Edition, calculate (a) the activation entropy, 4S, and (b) the activation enthalpy, 4H, for diffusion of carbon in bec iron. (Take v= 1077 73.) SOLUTION (a) The lattice parameter for bec iron is related to the atomic radius by as shown in Practice Problem 3.2 of Shackelford, Sth Edition Using Appendix 2 of Shackelford, Sth Edition, 4 (0.124 nm) = 0.286 nm. B As AG = OH - TAS, we can rewrite the model equation as: p= 2batvehS/® .- SH/R0, Comparing with the empirical diffusivity equation (Equation 5.13 in Shackelford, Sth mdition), we see that b= batvelS/R, 562 Table 5.2 gives, for carbon in bec iron, Di = 220% 10° m/s. Therefore, tse 2 Po ay = 1.61 x 10? AS = an (1.61 x 10°) = 7.306. Therefore, ds = 7.386 (8.314 3/(n01-K)) = 61.4 (3/K)/mol. (b) From the discussion relative to Equation 7-4, we note that From Table 5.2, AH = Q = 122 kJ/mol. 563 Materials engineers must be aware of the consequences of the chemical reactions between metals and their environments, as discussed in Chapter 19 of Shackelford, Sth Edition. These reactions include (a) direct oxidation by atmospheric gases and (b) various forms of electrochemical corrosion. We can better appreciate the nature of these various reactions by considering the associated thermodynamics. The nature of chemical degradation of ceramics and polymers is governed by similar thermodynamic principles. As indicated in Section 4, a chemical reaction is spontaneous if it is accompanied by a negative change in free energy. For the general reaction of a metal, M, with oxygen at a given temperature and oxygen pressure, aus E ose) = 40, (e-1) the change in Gibbs free energy, AG, is generally negative, as shown in Table 8-1. Consistent with our common experience, metals react with oxygen to some degree. The two exceptions in Table 8-1, silver and gold, are highly prized for the "noble" character. As with other processes discussed in Section 4,the AG for oxidation reactions are functions of the reaction temperature. In addition, the value of AG depends on the partial pressure of oxygen in the atmosphere. Some metals, such as palladium, can react with oxygen under a partial pressure of 1 atmosphere but will be unreactive (i.e., have a positive AG) under a modest vacuum. Corrosion is associated with aqueous environments. Consider the oxidation reaction for a metal, M,, going into aqueous solution: (8-2) where the ion in solution, M,"", has a valence of +n due to the liberation of n electrons. If an adjacent metal, Mz, in the aqueous environment is reduced by the electrons provided in Equation 8-2, the reaction will be (8-3) 564 Combining Equations 8-2 and 8-3 gives a net reaction: Meu = wt +m, tM) 7 (8-4) There is a negative value of AG for this spontaneous chemical reaction. The Gibbs free energy change is, in turn, equal to the electrical work done in the reversible transfer of electrons from metal M, to metal M2: As -nFV (8-5) electrical where n is the number of moles of electrons transferred per mole of M, ionized, F is Faraday's® constant (= 96,500 C/mol of electrons), and V is the voltage between the two metals. ‘The negative sign in Equation 8-5 is assigned to permit a positive voltage when the free energy change is nega- tive. In considering reactions such as Equation 8-4, the corrosion (i.e., ionization) of metal M, is spontaneous when AG is negative. Various electro- chemical cells associated with practical corrosion problems are illustrated in Sections 19.2 to 19.5 of Shackelford, 5th Edition “Michael Faraday (1791-1867), English chemist and physicist. Raised in poverty in London, Faraday's future lay in a career as a bookbinder until a chance opportunity for the bright and curious Faraday to apprentice to the eminent Humphry Davy. Faraday seized the opportunity and began one of the brilliant careers of nineteenth century science. He extended the research of predecessors and contemporaries Ampere, Coulomb, and Henry. In turn, James Clerk Maxwell was to provide an elegant mathematical extension of Faraday's concepts of field theory. (See Shackelford, Sth xdition, for related biographies.) 565 TABLE 8-1 Free Energy of Formation of Metal Oxides (per Gram-Atom of Oxygen) at 500 K and py = 1 atm.~ 2 Metal Oxide AG (kg) ca cao 578.5 Mg Mgo 547.5 AL A103 ~505.2 Ti Tos -423.6 Na Naz0 347.4 ce crea 341.6 an 2n0 298.5 a ih) 6 244.0 Fe Fe.03 232.3 co coo. 200.5 Mi Nio 193.0 cu cu.0 -131.9 ag Ag20 + 25 au Au203 + 44.0 "source: Data from J.H. Brophy, R.M. Rose, and J. Wulff, The Structure and Properties of Materials, Vol. 2: Thermodynamics of Structure, John Wiley & Sons, New York, 1964. 566 SAMPLE PROBLEM 8-1 What’ is the frée energy change associated with the reaction of aluminum with oxygen (p = 1 atm) at 500 K to form one mole of Al.0.? SOLUTION Relevant experimental data are given in Table 8-1 but on a basis of one gram-atom of oxygen. Considering the general reaction 2aL+30 = alo, it is clear that the formation of 1 mole of Alz03 requires 3 gram-atons of oxygen (actually 3/2 moles of diatomic 0,). Then, the AG for the above reaction will be three times the value given in Table 8-1: AG = 3 (-505.2 kJ) = -1515.6 kg. 567 SAMPLE PROBLEM 8-2 What is the free energy change associated with the corrosion of zn in the presence of Cu if the voltage between the two metals is measured to be 1.1 V? SOLUTION The periodic table (¢.g., Figure 2-2 in Shackelford, 5th Eéition) indicates that Zn is in group II B and will tend to produce a ivalent ion in solution: a = mnt +2. Using Equation 8-5, ac = -nFV =(2 mol)(96.5 % 10° C/mol)(1.1 V) 7212.3 x 10° Cv 7212.3 kJ. 568 SUMMARY ‘Thermodynamics plays a central role in materials science and engineer- ing. The first law of thermodynamics is a statement of the conservation of energy. The internal energy, U, of a system is a state variable. For a given process, AU is equal to the difference between the heat absorbed and ‘the work done (by the system). Relative to the first law, it is convenient 69/dt), and enthalpy, H (=U + pv). The second law states the limitations on conversion of heat to work. A state variable called entropy, 8, is introduced for the purpose of a formal statement of the second law. Specifically, we find that “any spontaneous to define heat capacity, C process is accompanied by an increase in total entropy." an equivalent "statistical" statement of the second law is that "a spontaneous process is one in which the system goes to a more probable state." A concept of importance to the description of solid solutions is the entropy of mixing, AS,. Defining the entropy of a system to be zero at absolute zero of temperature allows absolute values of entropy to be calculated at other temperatures and is taken as the third law of thermodynamics. Free energies are defined as “thermodynamic potentials," i.e., state variables that are minimized at equilibrium. For a process at constant temperature and pressure, the appropriate variable is the Gibbs free energy, G, defined as the enthalpy minus the product of temperature and entropy (HTS). The Gibbs free energy is the most generally useful thermodynamic potential for materials systems. For example, at atmospheric pressure, it is @ convenient indicator of the stable phases present at given temperatures and compositions. As such, the Gibbs free energy provides a theoretical basis for the phase rule and practical phase diagrams. The Gibbs free energy of surfaces (a type of planar defect) is closely related to surface tension, a mechanical quantity that can be experimentally determined in a variety of ways. The Gibbs free energy is also intimately related to kinetic rate equations. Careful inspection of this relationship indicates that an activation energy is an enthalpy term and that the pre-exponential constant contains an entropy term. As such, the Gibbs free energy provides a theoretical basis for the practical discussion of heat treatment. Finally, the nature of the oxidation and corrosion of metals by their environment is better appreciated by considering the associated thermodynamics. 569 activation energy adiabatic binary system Carnot cycle classical thermodynamics components conservation of energy contact angle corrosion cycle efficiency enthalpy enthalpy of mixing entropy entropy of mixing environmental reaction first law free energy Gibbs free energy Gibbs phase rule groove angle heat heat capacity heat treatment Helmholtz free energy ideal solution internal energy KEY WORDS 570 kinetics nonwetting oxidation partial wetting phase phase diagram phase transformation preexponential constant reversible second law sessile drop experiment state function state variable statistical thermodynamics Stirling's approximation surface surface energy surface tension system thermal grooving experiment ‘thermodynamic potential thermodynamics third law wetting work Brophy, J-H., R.M. Rose, and J. Wulff, The Structure and Properties of Materials, Vol.2: Thermodynamics ‘of Structure, John Wiley and Sons, Inc., New York, 1964. DeHoff, R.T., Thermodynamics in Materials Science, McGraw-Hill, New York, 1993. Holman, J.P., Thermodynamics, 4th Edition, McGraw-Hill, New York, 1988. Swalin, R.A., Thermodynamics of Solids, 2nd gdition, John Wiley and Sons, Inc., New York, 1972. sm

Vous aimerez peut-être aussi