You are on page 1of 30
Chapter 6 Thermodynamic Properties of Fluids The phase rule (Sec. 2.7) tells us that specification of a certain number of intensive properties of a system also fixes the values of all other intensive properties. However, the phase rule Provides no information about how values for these other properties may be obtained. Numerical values for thermodynamic properties are essential to the calculation of heat and work quantities for industrial processes. Consider, for example, the work requirement of a steady-state compressor designed to operate adiabatically and to raise the pressure of a gas from Py to P:. This work is given by Eq. (2.33), wherein the small kinetic- and potential- energy changes of the gas are neglected: W, = OH =H: — Hy Thus, the shaft work is simply 4H, the difference between initial and final enthalpy values of the gas. nr initial purpose in this chapter is to develop from the first and second laws the funda- scatal property relations which underlie the mathematical structure of thermodynamics. From these, we derive equations which allow calculation of enthalpy and entropy values from PVT and heat-capacity data. We then discuss the diagrams and tables by which property values are presented for convenient use. Finally, we develop generalized correlations which provide estimates of property values in the absence of complete experimental information. 6.1 PROPERTY RELATIONS FOR HOMOGENEOUS PHASES Equation (2.6), the first law for a closed system of m moles, may be written for the special case of a reversible process: (QU) = d Qe + d Wooy Equations (1.2) and (5.12) as applied to this process are: dWre = —PdinV) dQrey = T d(nS) 200 CHAPTER 6, Thermodynamic Properties of Fluids 6D where U, Sand V are molar values of the internal energy. entropy, and volume. ‘This equation, combining the first and second laws, is derived for the special case of a reversible process. However, it contains only properties of the system. Properties depend on state alone, and not on the kind of process that leads to the state. Therefore, Eq. (6.1) is not restricted in application to reversible processes. However, the restrictions placed on the nature of the system cannot be relaxed. Thus Eq. (6.1) applies to any process in a system of constant ‘mass that results in a differential change from one equilibrium state to another. The system may consist of a single phase (a homogeneous system), or it may be made up of several phases (a heterogeneous system): it may be chemically inert, or it may undergo chemical reaction. The only requirements are that the system be closed and that the change occur between equilibrium states. Alll of the primary thermodynamic properties— P, V, T, U, and S— are included in Eg. (6.1). Additional thermodynamic properties arise only by definition in relation to these Primary properties. The enthalpy, defined In Chap. 2 as a mater of convenience, is: HeU+PV @up ‘Two additional properties, also defined for convenience. are the Helmholtz energy and the Gibbs energy: AsU-TS (6.2) (63) Each of these defined properties leads directly to an equation like Eq. (6.1). Upon mul- tiplication by n, Eq. (2.11) may be differentiated to yield: dinH) = d(nU) + P d(aV) + (nV)dP ‘When dint!) is replaced by Eq. (6.1), this reduces to: d(aH) = T d(aS) + (nV)dP (6A) Similarly, Eq. (6.2) may be multiplied by 7, differentiated, and combined with Eq. (6.1): dna) In analogous fashion, Eqs. (6.3) and (6.4) combine to yield: d(nG) = (nV)dP — (nS\dT (6.6) ‘Equations (6.4) through (6.6) are subject to the same restrictions as Eq. (6.1). All are written for the entire mass of any closed system. ‘Our immediate application of these equations is to one mole (or to a unit mass) of a homogeneous fluid of constant composition. For this case, they simplify to: d(nV) — (aSdT (65) Obras prot 6.1. Property Relations for Homogeneous Phases 201 dU=TdS—PdV (67)|\dH=TdS+VdP (68) dA=—PdV-SdT (69)|dG=VdP-SdT (6.10) These fundamental property relations are general equations for a homogeneous fluid of constant composition. Another set of equations follows from Eqs. (6.7) through (6.10) by application of the criterion of exactness for a differential expression. If F = F(x, »), then the total differential of F is defined as: oF ara(Z axl, or (611) ae ) Ts er an er a tyar 2), Be ay ‘The order of differentiation in mixed second derivatives is immaterial, and these equations combine to give: aM an 6.12) (F),-(@), tsi When F is a function of x and y, the right side of Eq. (6.11) is an exact differential expression; because Eq. (6.12) must then be satisfied, it serves as the criterion of exactness. ‘The thermodynamic properties U, H, A, and G are known to be functions of the vari- ables on the right sides of Egs. (6.7) through (6.10); we may therefore write the relationship expressed by Eq. (6.12) for each of these equations. They are known as Maxwell's equations.! (,-@), «| @),-@),_ as -(#), (6.16) (i), Equations (6.7) through (6.10) are the basis not only for derivation of Maxwell's equa- tions but also of a large number of other equations relating thermodynamic properties. We develop here only a few expressions useful for evaluation of thermodynamic properties from experimental data. Their derivation requires application of Eqs. (6.7), (6.8), (6.15), and (6.16). 4 After James Clerk Maxwell (1831-1879), Scottish physicist. 22 (CHAPTER 6. Thermodynamic Properties of Fluids Enthalpy and Entropy as Functions of T and P ‘The most useful property relations for the enthalpy and entropy of a homogeneous phase result when these properties are expressed as functions of T and P. What we need to know is how Hi and S vary with temperature and pressure. This information is contained in the derivatives (8H/9T)p, (85/87), (@H/OP)y, and (88/3 Pr. Consider first the temperature derivatives, Equation (2.20) defines the heat capacity at constant pressure: aH (3), -¢ 220) Another expression for this quantity is obtained by division of Eq. (6.8) by dT and restriction of the result to constant P: (sr), (5) at), \aT) p ‘Combination of this equation with Eq. (2.20) gives: (F cr at), T ‘The pressure derivative of the entropy results directly from Eq. (6.16): as av (i), --(), on ‘The corresponding derivative for the enthalpy is found by division of Eq. (6.8) by dP and festriction to constant T: aH as (iF), = (Ge), +" As a result of Eq. (6.18) this becomes: an av (3),- v-r(z) (6.19) ‘The functional relations chosen here for H and S are H = H(T, P) and S = S(T, P). Whence, an aH d= (3), a7+ ap), a? ‘The partial derivatives are given by Eqs. (2.20) and (6.17) through (6.19): (6.17) au = cpar+[v-r(3) Je (6.20) aT) p. (621) ‘These are general equations relating the enthalpy and entropy of homogeneous fluids of con- ‘stant composition to temperature and pressure. 6.1. Property Relations for Homogeneous Phases 203 Internal Energy as a Function of P Internal energy is given by Eq. (2.11) as U = H ~ PV. Differentiation yields: ( (const T) (6.49) The compressibility factor is defined as Z = PV/RT; values of Z and of (8Z/AT)p may be calculated from experimental PVT data. with the two integrals in Eqs. (6.46), (6.48), und (6.49) evaluated by numerical or graphical methods. Alternatively, the two integrals may be evaluated analytically when Z is expressed as a function of T and P by a volume-explicit equation of state. This direct connection with experiment allows evaluation of the residual properties 7" and S*, which are essential to the practical application of thermodynamics. Residual Properties in the Zero-Presure Limit ‘The constant J, omitted from Eqs. (6.46), (6.48), and (6.49), is the value of G*/RT in the limit as P — 0. Its origin becomes clear from a general treatment of residual properties in this limit. Because a gas becomes ideal as P + 0 (in the sense that Z — 1), one might ‘suppose that in this limit all residual properties are zero, This is not in general true, as is easily demonstrated through consideration of the residual volume. Written for V¥ in the limit of zero pressure Eq. (6.41) becomes, lim V® = tim ¥ — lim v'F Po PO Pos Both terms on the right are infinite, and their difference is indeterminate, Experimental insight is provided by Eq. (6.40): z-1 az i Re i —)= it — AS” = RF fn, ( P ) ar jin, (7F) ‘Thus, V®/RT in the limit as P —> 0 ata given T equals the slope of the Z vs. P isotherm at P =0. Figure 3.9 shows clearly that these values are finite, and not in general zero. For the intemal energy, U*® = U — U's. Because U'* is a function of T only, a plot of U's ys. P for a given T is 2 horizontal line extending to P = 0. For a real gas with 6.2. Residual 2 finite intermolecular forces, an isothermal expansion to P — 0 results in an increase in U, because the molecules move apart against the forces of intermolecular attraction. Expansion to P =0(V = oe) reduces these forces to zero, exactly as in an ideal gas, and therefore at all temperatures, lim U =U"* and lim U* =0 pa, pao From the definition of enthalpy, oe em tS z Bit pts payers Because both terms on the right are zero, lim H = 0 for all temperatures. For the Gibbs energy, by Eq. (6.37): G)_v «(Z) =ppiP (om T) For an ideal gas, V = V'* = RT/P. and this becomes: Gt) _4P “(G)= “p omn Integration from P =O to pressure P yields: Ge Gis ? aP Gs o-(%),.*f = (),.*P +” (const T) For finite values of G'#/RT at P > 0, we must have lim (G'"/RT) = —o0, Because thi true for G as well, we conclude that GF G Git ny ee = Par Ber Te Thus G*/RT (and of course G®) is, like V®, indeterminate im the limit as P —> 0. In this case however, no experimental means exists for finding the limiting value. However, we have no reason to presume it is zero, and therefore regard it like limp_.o V* as finite, and not in general zero. Equation (6.44) provides an opportunity for further analysis. We write it for the limiting case of P = 0: i) _ fae /RT) RF} pg aT pao ‘Asalready shown, H®(P = 0) = 0. and therefore the preceding derivative is zero. Asa result, (i RT} poo where J is a constant, independent of T. 212 CHAPTER 6. Thermodynamic Properties of Fluids Enthalpy and Entropy from Residual Properties Apptied to the enthalpy and entropy, Eq. (6.41) is written: AH=H®+H* Sasa s® ‘Thus, H and $ follow from ideal-gas and residual properties by simple addition. General expressions for H’® and S' are found by integration of Eqs. (6.23) and (6.24) from an ideal- ‘gas state at reference conditions Ti and Pp to the ideal-gas state at T and P=* To igdT P ait = es + f Ma raf ce— - Rin— a+ J + fcr 7 Substitution into the preceding equations gives: rT waits [ crar+H® (6.50) S= sof cp - rine + st 51) Recall (Secs. 4.1 and 5.5) that for purposes of computation the integrals in Eqs. (6.50) and (6.51) are represented by: * f Cif aT = R x ICPH(TO,T:A.B.C,D) f cet = = R x ICPS(T0,7:A.B.C.0) Equations ($0) and St) have aterative form when the integra replace by equivalent terms that include the mean heat capacities introduced in Secs. 4.1 and 5.5: HH + (Ci) (7 - Ty) + H* (6.52) S=SHICS gin — Rint + 5" (6.53) In Eqs. (6.50) through (6.53), H* and S* are given by Eqs. (6.46) and (6.48). Again, for ‘computational purposes, the mean heal capacities are represented by: (CE = & x MCPH(TO,T:A,B,C,D) (Ci)g = R x MCPS{T0,T;A,8,C,D) Applications of thermodynamics require only differences in enthalpy and entropy. The reference-state conditions To and Py are therefore selected for convenience, and values are assigned to #,* and S;* arbitrarily. The only data needed for application of Eqs. (6.52) and (6.53) are ideal-gas heat capacities and PVT data. Once V, H, and § are known at given conditions of 7 and P. the other thermodynamic properties follow from defining equations. ‘Thermodynamic properties for emt compere is enh an ou mm sre 74 Fetal. 1. Kabo, KN. Marsh, G.N. Roganoy, amd R. C. Wilbolt. Thermodynamics of Ongienic Compounds in the Gas Stote, Theems- dynamics Research Center, Texas A & M Univ. Systeas. College Station, Teas, 199, 6.2. Residual Properties 213 The true worth of the equations for ideal gases is now evident. They are important because they provide a convenient base for the calcu- lation of real-gas properties. Residual properties have validity for both gases and liquids. However, the advantage of Eqs. (6.50) and (6.51) in application to gases is that H'® and S®, the terms which contain all the complex calculations, are residuals that are usually small. They act as corrections to the major terms, H'® and S"". For liquids, this advantage is largely lost, because H* and S® must include the large enthalpy and entropy changes of vaporization. Property changes of liquids ‘are usually calculated by integrated forms of Eqs. (6.28) and (6.29), as illustrated in Ex. 6.1. Example 6.3 Calculate the enthalpy and entropy of saturated isobutane vapor at 360 K from the following information: 1. Table 6.1 gives compressibility-factor data (values of Z) for isobutane vapor. 2. The vapor pressure of isobutane at 360 K is 15.41 bar. 3. Set Hi = 18,115.0 J mol~! and Sif = 295.976 J mol! K-! for the ideal-gas reference state at 300 K and 1 bar. [These values are in accord with the bases adopted by R. D. Goodwin and W. M. Haynes, Nat. Bur. Stand. (U.S.), Tech, Note 1051, 1982] 4. The ideal-gas heat capacity of isobutane vapor at temperatures of interest is: CE/R = 1.7765 + 33.037 x 10° T (7 / Ky Table 6.1: Compressibility Factors Z for Isobutane Phar MOK 350K 360K 370K 380K 0.10 099700 0.99719 0.99737 0.99753 0.99767 0.50 0.98745 0.98830 0.98907 0.98977 0.99040 = 0.95895 0.96206 0.96483 0.96730 0.96953 4 0.92422 0.93069 0.93635 0.94132 0.94574 6 0.88742 0.89816 0.90734 0.91529 0.92223 8 O84575 0.86218 0.87586 0.88745 0.89743 10 0.79659 0.82117 0.84077 0.85695 0.87061 . - ©.77310 0.80103 0.82315 0.84134 0.75506 0.78531 0.80923 0.71727 1541 Solution 6.3 Calculation of H® and S® at 360 K and 15.41 bar by application of Eqs. (6.46) and (6.48) requires evaluation of two integrals: . dP F dP f (%),F [oe-v¥ Graphical i requires simple plots of (82/37) p/P and (Z—1)/P vs. P. Values of (Z — 1)/ P are found from the compressibility-factor data at 360 K. The quantity (82/9T}p/P requires evaluation of the partial derivative (3Z/8T )p. given by the slope of a plot of Z vs. T at constant pressure. For this purpose, sep- arate plots are made of Z vs. T for each pressure at which compressibility-factor data are given, and a slope is determined at 360 K for each curve (for example, by construction of a tangent line at 360 K). Data for the required plots are shown in Table 6.2. ‘Table 6.2: Values of the Integrands Required in Ex. 6.3 ‘Values in parentheses are by extrapolation. Pfoar ((8Z/T)p/PI x 10K" bart 0 O10 050 2 4 6 8 10 12 14 2.432 1541 (2.720) ‘The values of the two integrals are: PF az\ dP ai e dP [ (=), p= 2637 x 10K [2 v$ = -02586 By Eq. (646), Gp = —(360)(26.37 x 10-*) = -0.9493 sf By Eq. (6.48), > = -0.9493 — (—0.2596) = -0.6897 CHAPTER 6, Thermodynamic Properties of Fluids Residual ies by Equations of State 215 For R = 8.314 J mol! K~', = (—0.9493)(8.314)G60) = -2,841.3 J mor S® = (—0,6897)(8.314) = -5.734J mot“! K-" ‘Values of the integrals in Eqs. (6.50) and (6.51) are: 8.314 x ICPH(300,360;1.7765,33.037E-3,0.0,0.0) = 6,324.8 J mot! 8.314 x ICPS(300,360;1.7765,33.037E-3,0.0,0.0) = 19.174) mol! K-! Substitution of numerical values into Eqs. (6.50) and (6.51) yields: A = 18,115.0 + 6324.8 — 2,841.3 = 21,598.5 Jmol J = 295.976 + 19.174 ~ 8.314 In 15.41 — 5.734 = 286,676 J mol! K~* Although calculations are here carried out for just one state, enthalpies and entropies can be evaluated for any number of states, given adequate data. After having completed a set of calculations. one is not irrevocably committed to the particular values of H* and Si* initially assigned. The scale of values for either the enthalpy or the entropy can be shifted by addition of a constant to all values. In this way one can give arbitrary values to H and S for some particular state so as to make the scales convenient for one purpose or another, The calculation of thermodynamic properties is an exacting task, seldom required of an engineer. However, engineers do make practical use of thermodynamic properties, and an understanding of calculational methods should suggest that some uncertainty is associated with every property value. Inaccuracy derives partly from experimental error in the data; they are also frequently incomplete and must be extended by interpolation and extrapolation. Moreover, even with reliable PVT data, a loss of accuracy occurs in the differentiation process required in the calculation of derived properties. Thus data of a high order of accuracy are required to produce enthalpy and entropy values suitable for engineering calculations. 6.3 RESIDUAL PROPERTIES BY EQUATIONS OF STATE An attractive alternative to the numerical evaluation of integrals in Eqs. (6.46) and (6.48) is their analytical evaluation by equations of state. This requires an equation which can be di- rectly solved for Z (or V) as a function of P at constant T. Such an equation of state is said to be volume explicit, and the only example presented in Chap. 3 is the virial expansion in P. The other equations of state are pressure explicit; i.e.. they can be solved for Z (ot P) as functions of V at constant T. They are not suitable for direct use with Eqs. (6.46) and (6.48). The virial ‘expansion in V and all cubic equations of state are pressure explicit,* and their use for evalua- tion of residual properties requires the reformulation of Eqs. (6.46), (6.48), and (6.49). In what “The ideal-gas equation is both pressure and volume explicit, 216 CHAPTER 6. Thermodynamic Properties of Fluids follows, we treat the calculation of residual properties for gases and vapors through use of the virial equations and cubic equations of state. Residual Properties from the Virial Equations of State Equation (3.38), the two-term virial equation, gives Z — 1 = BP/RT. Substitution in Eq. (6.49) reduces it to: . G BP RT RT 6) He ‘a(G*/RT) P\(/\dB B mean Fool] (BN (G8-#) a* B 4B or 7s (F = 3) (6.55) Substitution of Eqs. (6.54) and (6.55) into Eq. (6.47) yields: sf PdB kar (6.56) Evaluation of residual enthalpies and residual entropies by Eqs. (6.55) and (6.56) is straight- forward for given values of T, P, and composition, provided onc has sufficient data to evaluate B and dB/dT. The range of applicability of these equations is the same as for Eq. (3.38), as discussed in Sec. 3.4. Equations (6.46), (6.48), and (6.49) are incompatible with pressure-explicit equations of state, and nmst be transformed to make V (or molar density p) the variable of integration. In application p is a more convenient variable than V, and the equation, PV = ZRT. is written in the alternative form, P=ZpRT (6.57) Differentiation gives: dP = RT(Zdp+pdZ) (const T) Tn combination with Eq. (6.57), this equation is recast: (658) where the integral is evaluated at constant 7. Note also that 9 > 0 when P — 0, ‘Solving Eq. (6.42) for its final term and substituting for V* by Eq. (6.40) yields: ap (at Mar = a-v$ a(S ) 6.3. Residual ie ions of State 217 Division by dT and restriction to constant p gives: B® Z- - ee") ° ar od, RPP Differentiation of Eq. (6.57) provides the first derivative on the right, and differentiation of Eq. (6.58) provides the second. Substitution leads to: He Z\ dp moot GF +e oo ‘The residual entropy is found from Eq. (6.47): =wz- rf ) 2 -fia- v% (6.40) The three-term virial equation is the simplest pressure-explicit equation of state: Z-1=Bp+Cp* G.40) Substitution into Eqs. (6.58) through (6,60) leads to: OF 289 + 3cy?—t02z (661) BERG =Ihz- r{(2+ Borg (E+) e'] (6.63) Application of these equations, useful for gases up to moderate pressures, requires dats for both the second and third virial coefficients. Residual Properties by Cubic Equations of State Results of some generality follow from application of the generic cubic equation of state: RT a(t) Vb (Vieb{¥ +06) Derivations with this equation are much more convenient when it is recast to yield Z with density p as the independent variable. We therefore divide Eq. (3.42) through by pRT and substitute V = 1/p, With q given by Eq. (3.51). the result after some alegbraic reduction is: a + epbyl + op) P= GA2) ze 218 CHAPTER 6. Thermodynamic Properties of Fluids ‘The two quantities needed for evaluation of the integrals in Eqs. (6.58) through (6.60), Z — | and (92/47), are readily found from this equation: z-1= se (6.64) — pb “TF epbyl Fepb) (5), -- Gt) arcatirrem at), ~~ \at ) OF epbyl ¢ enh) ‘The integrals of Eqs. (6.58) through (6.60) are now evaluated as follows: fie-ve-[ pb dpb) _ finite Is ? T= pb pb "J, UF Eph + op6) f 2) dp __ dq f” ap) lo (i pe aT Sy (+e pb +apb) ‘These two equations simplify to: dp "(z) dp ___ dq = ple ania = p= a2) ey fe nF In] — pb) gt f aT}, Pp aT a - dipb) where by definition, =f = Wrephyn sorb (const T) ‘The generic equation of state presents two cases for the evaluation of this integral: Case: «x0 t= ae) (6.65a) ane l+epb Application of this and subsequent equations is simpler when p is eliminated in favor of Z. By ‘Eq. (3.50) and the definition of Z: bP P La Aaa Zs— whence zu (6.656) ‘Case Hr 6 = se ze =o aR ‘The van der Waals equation is the only one considered here to which Case II applies. and this equation then reduces to f= 6/Z. With evaluation of the integrals, Eqs. (6.58) through (6.60) reduce to: G® ar 727 tml eb)Z at (6.664) as por direit 6.3_Residual c ions of State 219 Gr or pr 7 2—'-WZ-B)—al (6.666) He dq Frnt ter (B)raz-144, RT s* dq and Gnz-o+ (a+ St): The quantity 7;(dq /dT,) is readily found from Eq, (3.54): dq _[dina(T;) nae-| dint, i] ‘Substitution for this quantity in the preceding two equations yields: H® dina(7,) wz-1+[ 7a ~]er (6.67) s =z a+ Anette 1 (6.68) i war ) Preliminary to application of these equations one must find Z by solution of Eq. (3.52) for a cporpearee (3.36) for a liquid phase. Example 6.4 Find values for the residual enthalpy H* and the residual entropy S* for n-butane gas at 500 K and 60 bar as given by the Redlich/Kwong equation. Solution 6.4 0 asa 7 1176 P= 5795 = 137 By Eq. (3.53), with © for the Redlich/Kwong equation from Table 3.1, p. 98, P, _ (0.08664)(1.317) Baar 1.176 With values for V and 2. and with the expression a(T,) = T;”"? from Table 3.1. Eq. (3.54) yields: = 0.09703 _ Walt) 0.42748 . = or, ~ W0seC.17H sam 20 CHAPTER 6. ties of Fluids ‘Substitution of 8.9, € = 0, and a = 1 into Eq. (3.52) reduces it to: Z — 0.09703 Z=1+0.09703 — (G.8689)(0.09703) 7 oo703) Solution of this equation yields Z = 0.6850. Then: ren2i? = 0.13247 With Ina(7,) = ~4 In T,. d Ina(T,)/d In T, = —}. Then Eqs. (6.67) and (6.68) become: ie a = 0.6850 — 1 + (-0.5 — 1)(3.8689)(0.13247) = - 1.0838 sk Sz = 010.6850 — 0.09703) — (0.5(3.8689)(0.13247) = 0.78735 Whence, HH = (8.314)(500)(~1.0838) = —4,505 J mol! S® = (8.314)(—0.78735) = —6.546 J mot! K-! These results are compared with those of other calculations in Table 6.3. ‘Table 6.3: Values for Z, H*, and S* for n-Butane at 500 K and 50 bar H®/}mot™'— S*/) mol“'K™ -3.937 -5.424 4,505 6.546 4,824 —7413 4,988 —7A 4,966 —7.632 4.760 7.170 * Described in Sec. 6.7. ¥ Values derived from numbers in Table 2-240, p. 2-223, Chemical Engineers’ Handbook, 7th ed., Don Green (ed.). McGraw-Hill, New York, 1997. 6.4 TWO-PHASE SYSTEMS ‘The curves shown on the PT diagram of Fig. 3.1 represent phase boundaries for a pure sub- stance. A phase transition at constant temperature and pressure occurs whenever one of these 6.4, Two-Phase Systems 21 curves is crossed, and as a result the molar or specific values of the extensive thermodynamic properties change abruptly. Thus the molar or specific volume of a saturated liquid is very dif ferent from the molar or specific volume of saturated vapor at the same T and P. This is true as well for internal energy, enthalpy, and entropy. The exception is the molar or specific Gibbs energy, which for a pure species does not change during a phase transition such as melting, vaporization, or sublimation. Consider a pure liquid in equilibrium with its vapor in a pis- ton/cylinder arrangement at temperature 7 and the corresponding vapor pressure P *. When a differential amount of liquid is caused to evaporate at constant T and P, Eq. (6.6) applied to the process reduces to d(nG) = 0. Because the number of moles n is constant. dG = 0, and this requires the molar (or specific) Gibbs energy of the vapor to be identical with that of the liquid. More generally, for two phases o and B of a pure species coexisting at equilibrium, =a (6.69) where G* and G* are the molar or specific Gibbs energies of the individual phases. ‘The Clapeyron equation, first introduced in Sec. 4.2, follows from this equality. If the temperature of a two-phase system is changed, then the pressure must also change in accord with the relation between vapor pressure and temperature if the two phases continue to coexist in equilibrium. Because Eq. (6.69) applies throughout this change, dG" = dG" ‘Substituting expressions for dG™ and dG? as given by Eq. (6.10) yields: ve dP™ — dT = VP dP™ ~SPaT which upon rearrangement becomes: ‘The entropy change AS and the volume change AV are changes which occur when a unit amount of a pure chemical species is transferred from phase a to phase f at the equilibrium T and P. Integration of Eq. (6.8) for this change yields the latent heat of phase transition: OH” =TAS® (6.70) ‘Thus, AS = SH /T. and substitution in the preceding equation gives: dp™ AH? dr TAve (67) which is the Clapeyron equation. For the particularly important case of phase transition from liquid ! to vapor v, Eq. (6.71) is writen: , dP“ ant ig * Fave (6.72) Obra: tegidas por direitos de < 2 CHAPTER 6, Thermodynamic Properties of Fluids But ave = SZ az where AZ" is the compressibility-factor change of vaporization. Combination of the last 1wo equations gives, on rearrangement: dinP™ AH" ar = Rraz (6.73) dinp™ Ani aT) RAZ Equations (6.72) through (6.74) are equivalent, exact forms of the Clapeyron equation for (6.74) Example 6.5 ‘The Clapeyron equation for vaporization may be simplified by introduction of reason- able approximations, namely, that the vapor phase is an ideal gas and that the molar volume of the liquid is negligible compared with the molar volume of the vapor. How do these assumptions alter the Clapeyron equation? Solution 6.5 ‘The assumptions made are expressed by: ee yt AT fe av’ wv’ “7a or Az =I ei te —ptinP™ Equation (6.74) yields: AH Rah ‘This approximate equation. known as the Clausius/Clapeyron equation, relates the latent heat of vaporization directly to the vapor-pressure curve. Specifically, it shows that AH'* is proportional to the slope of a plot of In P™ vs. 1/T. Such plots of experimental data produce lines for many substances that are nearly straight. According to the Clausius/Clapeyron equation, this implies that 4H'" is almost constunt, virtually independent of T. This is not true: AH! decreases monotonically with increasing temperature from the triple point to the critical point, where it becomes zero. The assumptions on which the Clausius/Clapeyron ‘equation are based have approximate vatidity only at low pressures. 6.4, Two-Phase Systems 23 Temperature Dependence of the Vapor Pressure of Liquids ‘The Clapeyron equation is an exact thermodynamic relation, providing a vital connection be- tween the properties of different phases. When applied to the calculation of latent heats of vaporization, its use presupposes knowledge of the vapor pressure-vs.-temperature relation. Because thermodynamics imposes no mode! of material behavior, cither in general or for par- ticular species, such relations are empirical. As noted in Ex. 6,5. a plot of In P“ vs. 1/T generally yields a line that is nearly straight: B ae InP A Tr (6.75) where A and B are constants for a given species. This equation gives a rough approximation of the vapor-pressure relation for the entire temperature range from the triple point to the critical point. Moreover, it provides an excellent basis for interpolation between reasonably spaced values of T. ‘The Antoine equation, which is more satisfactory for general use, has the form: a T+C A principal advantage of this equation is that values of the constants A, B, and C are readily available for a large number of species.> Each set of constants is valid for a specified tempera- ture range. and should not be used much outside of that range. Values of Antoine constants for selected substances are given in Table B.2 of App. B. ‘The accurate representation of vapor-pressure data over a wide temperature range re- ‘quires an equation of greater complexity. The Wagner equation is one of the best available; it expresses the reduced vapor pressure as a function of reduced temperature: 33034 Dr5 taps = AEBS CO + Det «om where r=1-f and A, B,C, and D are constants. Values of the constants either for this equation or for Eg. (6.76) are given by Reid, Prausnitz, and Poling® for many species. Corresponding-States Correlations for Vapor Pressure A number of corresponding-states correlations are available for the vapor pressure of non- polar, non-associating liquids. One of the simplest is that of Lee and Kesler.’ {tis a Pitzer-type correlation, of the form: InP™=A-— 6.76) In PS4T,) = In PPT,) +eln PT) (6.78) 53. Ohe, Computer Aided Dato Book of Vapor Pressure, Data Book Publishing Co Tokyo, 1976: T. Boublik, V. Bried, and E. Haka, The Vapor Pressures of Pure Substances, Elsevier, Amstentam, 1984. R.C. Reid, J. M, Prausnitz, and B. E. Poling, The Properties of Gases and Liguids, 4th ed, App. A, McGraw-Hill, 1987. 7B.L Lee and M.G. Kester, ACRE J., vol, 24, pp. 310—S27, 1975, 224 CHAPTER 6. Thermodynamic Properties of Fluids where In PO(T,) = 5.92714 — — 1.28862 In T, + 0.1693477° (6.79) 6.09648 T 15.6875 tr ‘Lee and Kesler recommend that the value of w used with Eq. (6.78) be found from the correla- tion by requiring that it reproduce the normal boiling point. In other words, w for a particular substance is determined from: In P}(7,) = 15.2518 — — 13.4721 In T, + 0.435778 (6.80) In P* — in POT,,) InPNT,,) where T,, is the reduced normal boiling point and P,%* is the reduced vapor pressure corre- sponding to 1 standard atmosphere (1.01325 bar). (681) Example 6.6 Determine the vapor pressure (in kPa) for liquid n-hexane at 0, 30, 60, and 90°C: (a) With constants from App. B.2. (b) From the Lee/Kesler correlation for P,. Solution 6.6 (a) With constants from App. B.2, the Antoine equation for n-hexane is: 2696.04 a - —— In P™/kPa = 13.8193 iP@+ 24317 Substitution of temperatures yields the values of P =" under the heading “Antoine” in the table below. We take these to be equivalent to good experimental values. (6) We first determine « from the Lee/Kesler correlation. At the normal boiling point of n-hexane (Table B.1), _ M19 wx _ 1.01325 Tre = Foqg = 0879 and an = Ss Application of Eq. (6.81) then gives the value of w for use with the Lee/Kesler correlation: @ = 0.298. With this value, the correlation produces the P =* values shown in the table. The average difference from the Antoine values is about 1.5%. = 0.03350 1PC PP /kPa P/kPa (Antoine) (Lee/Kesier) 0 6.052 5.835 oO 7646 76.12 rec PS kPa P“kPa (Antoine) (Lee/Kesler) 30 24,98 24.49 90 189.0 190.0 65. Thermodynamic Diagrams 25 Two-Phase Liquid/ Vapor Systems ‘When a system consists of saturated-liquid and saturated-vapor phases coexisting in equilib- rium, the total value of any extensive property of the two-phase system is the sum of the total Properties of the phases. Written for the volume, this relation is: alVi tnt? where V is the molar volume for a system containing a total number of moles n =n! +n°. Division by mt gives: nv Ven eatVe ype x! and represent the mass fractions of the total system that are liquid and vapor With =l-x". V=(l-x")¥4x°V" In this equation the properties V, V', and V* may be cither molar or unit-mass values. The mass or molar fraction of the system that is vapor x” is called the quality. Analogous equations can be written for the other extensive thermodynamic properties. All of these relations are represented by the generic equation: Ma (1 ~xM! +2°M" (6.82a) where M represents V,U, H. S.etc. An alternative form is sometimes useful: M=M!'+x°aM"* (6.82) 6.5 THERMODYNAMIC DIAGRAMS A thermodynamic diagram is a graph showing for « particular substance a set of properties, eg. T, P. V. H. and S, The most common diagrams are: TS, PH (usually In P vs. H), and HS (called a Mollier diagram). ‘The designations refer to the variables chosen for the coordinates. Other diagrams are possible, but are scldom used. Figures 6.2 through 6.4 show the general features of these diagrams. Though based on data for water, their gencral character is similar for all substances, The two-phase states, represented by lines on the PT diagram of Fig. 3.1, lie over areas in these diagrams, and the triple point of Fig. 3.1 becomes a line, Lines of constant quality in a liquid/vapor region, provide directly two-phase property values. The critical point is identified by the letter C, and the solid curve passing through it represents states of saturated tiquid (to the left of C) and of saturated vapor (to the right of C). The Mollier diagram (Fig. 6.4) does not asually include volume data. In the vapor or gas region. lines for constant temperature and constant superheat appear. Superheat is a term denoting the difference between the actual temperature and the saturation temperature at the same pressure. Thermodynamic diagrams included in this book are the PH diagrams for methane and tetrafluoroethane in App. G, and the Mollier diagram for steam on the inside of the back cover. Paths of processes are easily traced on a thermodynamic diagram. For example, the boiler of a steam power plant has liquid water as feed at a temperature below its boiling poiat, Figure 6.2: PH diagram. ‘and superheated steam as product. Thus, water is heated at constant P to its saturation temper- ature (line I~2 in Figs. 6.2 and 6.3), vaporized at constant T and P (line 2~3), and superheated at constant P (line 3-4). On a PH diagram (Fig. 6.2) the whole process is represented by a horizontal line corresponding to the boiler pressure. The same proess is shown on the TS diagram of Fig. 6.3. The compressibility of a liquid is small for temperatures well below T., ‘and liquid-phase properties change very slowly with pressure. The constant-P lines on this diagram for the liquid region therefore lie very close together, and line 1-2 nearly coincides with the saturated-liquid curve. The isentropic path of the fluid in a reversible adiabatic turbine ‘of compressor is represented on both the T'S and H § (Mollier) diagrams by a vertical line from the initial to the final pressure. 6.6 TABLES OF THERMODYNAMIC PROPERTIES Tn many instances thermodynamic properties are tabulated. This has the advantage that data can be presented more precisely than in diagrams, but the need for interpolation is introduced. Thermodynamic tables for saturated steam from its normal freezing point to the critical point and for superheated steam over a substantial pressure range, in both S] and in English units, appear in App. F. Values are given at intervals close enough that linear interpolation is satisfactory.® The first table for each system of units shows the equilibrium properties of satu- rated liquid and saturated vapor at even increments of temperature. The enthalpy and entropy are arbitrarily assigned values of zero for the saturated-liquid state at the triple point. The sec- ond table is for the gas region, and gives properties of superheated steam at temperatures higher "Procedares For linear interpolation arc shown at the beginning of App. F Obras 6.6. Tables of Thermodynamic Properties 21 Figure 6.4: Mollier diagram. than the saturation temperature fora given pressure. Volume (¥), internal energy (U), enthalpy (H), and entropy (S) are tabulated as functions of pressure at various temperatures. The steam tables re the most thorough compilation of properties for any single material. However tables are available for a number of other substances. Example 6.7 ‘Superheated steam originally at P; and 7; expands through a nozzle to an exhaust Pressure P;. Assuming the process is reversible and adiabatic, determine the down- ‘Stream state of the steam and AH for the following conditions: (2) Pi) = 1,000 kPa, 1 = 250°C, and P; = 200 kPa. () Py = 150(psia), 11 = S00(°F), and P; = 50(esia). “Data tor many common chemicals are given By RH. Peery ated D. Green. Perry's Chemical Exgineers" Hand- book. Tth ed., Sec. 2, McGraw-Hill, New York, 1996, Sce also N. B. Vargafllk. Handbook of Physical Properties of Liquids and Gases, 24 ¢8., Hemisphere Publishing Corp.. Washington, DC, 1975, Data for refrigerants appear in the ASHRAE Handbook: Fundamentals, American Society of Heating. Refrigerating. and Air-Conditioning Engineers, nc. Attanta, 1993, Solution 6.7 Because the process is both reversible and adiabatic, there is no change in the entropy of the steam, (a) For the initial temperature of 250°C at 1,000 kPa, no entries appear in the SI tables for superheated steam, Interpolation between values for 240°C and 260°C yields, at 1000 kPa: Hy = 2.9429 ke 5S; = 6.9252 kg! K™? For the final state at 200 kPa, Sz = 5; = 6.9252 I kg! K- Because the entropy of saturated vapor at 200 kPa is greater than Sp, the final state is in the two-phase liquid/vapor region. Thus /2 is the saturation temperature at 200 kPa, given in the SI superheat tables as t = 120.23°C. Equation (6,82) applied to the entropy becomes: S2= (1 -x3)Sh + 355$ ‘Whence, 6.9252 = 1.5301(1 — xf) + 7.12683 where the values 1.5301 and 7.1268 ure entropies of saturated liquid and saturated vapor at 200 kPa. Solving, xf = 0.9640 On a mass basis, the mixture is 96.40% vapor and 3.60% liquid. Its enthalpy is ‘obtained by further application of Eq. (6.82a): Hz = (0.0360)(504.7) + (0.9640)(2,706.3) = 2,627.0 kJ kg! Finally, AH = Hy ~ Hy = 2,627.0 — 2,942.9 = -315.9 kJ ke! (b) For the initial state at 150(psia) and S00(*F) data from Table F:4 for super- heated steam in English units provide: Ay = 1,274.3 (Bua\lba)! $1 = 1.6602(Btu) (Ibm) "(RY In the final state at SO

You might also like