<<

Chapter 5 Thermodynamic Properties of Real

It has already been demonstrated through the first and second laws, that the and interactions between a system and its surroundings may be related to the state variables such as internal , and . So far we have illustrated the calculations of energy and entropy primarily for pure (component) ideal systems. However, in practice this is an exception rather than a rule as one has to deal with not only removed from state, but also with and . In addition, mixtures rather than pure components are far more common in chemical process plants. Therefore, computation of work and heat interactions for system comprised of real fluids requires more complex thermodynamic formulations. This chapter is devoted to development of such relations that can help calculate energy requirements for given changes of state for real systems. As in the case of ideal gases the goal is to correlate the energy and entropy changes for real fluids in terms of their volumetric and other easily measurable macroscopic properties.

5.1 Thermodynamic Property Relations for Single Systems Apart from and enthalpy, two other ones that are particularly useful in depiction of thermodynamic equilibrium are (A) and (G). We defer expanding upon the concept of these two types of to chapter 6; however, we state their definition at this point as they are instrumental in the development of property correlations for real fluids. • Specific Helmholtz free energy: A= U − TS ..(5.1) • Specific Gibbs free energy:G= H − TS ..(5.2) For a reversible process in a the first law gives:

dU= dQ + dW

Or:

dU= TdS − PdV ..(5.3)

Using H= U + PV and taking a total differential of both sides:

1 | Page

dH=++ dU PdV VdP ..(5.4)

Putting eqn. 5.3 in 5.4 we get:

dH= TdS + VdP ..(5.5)

In the same manner as above one may easily show that the following two relations obtain: dA=−− SdT PdV ..(5.6)

dG= VdP − SdT ..(5.7)

Equations 5.3 to 5.7 comprise the fundamental energy relations for thermodynamic systems where there is a single phase with constant composition. In principle, they may be integrated to compute the energy changes for a system transiting from one equilibrium state to another.

5.2 All the four types of energy relations above satisfy the mathematical condition of being continuous variables, as they are themselves functions of state variables. One can thus apply of the criterion of for these functions. For a function of the form P= PXY(,) one can write the following total differential:

∂∂PP  dP=+=+ dX  dY MdX NdY ..(5.8) ∂∂XYYX  ∂∂PP  Where: M=  and N =  ..(5.9) ∂∂XYYX  ∂M ∂22 P ∂∂ NP Further, = and = ..(5.10) ∂YXY ∂∂ YX ∂ X ∂∂ XY ∂∂MN   It follows:  =  ..(5.11) ∂∂YX XY  Applying eqn. 5.11 to 5.3, 5.5, 5.6 and 5.7 one may derive the following relationships termed Maxwell relations: ∂∂TP  = −  ..(5.12) ∂∂VSSV  ∂∂TV    =  ..(5.13) ∂∂PS SP  2 | Page

∂∂PS    =  ..(5.14) ∂∂TV VT  ∂∂VS  = −  ..(5.15) ∂∂TPPT 

5.4 Relations for Enthalpy, Entropy and Internal Energy One may conveniently employ the general energy relations and Maxwell equations to obtain expressions for change in enthalpy and entropy and internal energy for any process, which in turn may be used for computing the associated heat and work interactions. Let H= HTP(, )

∂∂HH  Then: dH= dT +  dP ∂∂TPPT 

∂H But = CP ∂T P

∂H Thus: dH= CP dT +  dP ..(5.16) ∂P T

∂∂HS   Using dH=+⇒ TdS VdP   =TV  + ..(5.17) ∂∂PP TT 

∂∂SV  From Maxwell relations as in eqn. 5.15:= −  ..(5.18) ∂∂PTTP 

Thus using eqns. 5.17 and 5.18 in 5.16 we get:

∂V dH= CP dT +− V T dP ..(5.20) ∂T P

In the same manner starting from the general function: U= UTV(, )and S= ST (, P ) and applying appropriate Maxwell relations one may derive the following general expressions for differential changes in internal energy and entropy.

∂P dU=+− CV dT T P dV ..(5.21) ∂T V

3 | Page

dT∂ V dS= CP − dP ..(5.22) TT∂ P

dT∂ P Or, alternately: dS= CV + dV ..(5.23) TT∂ V

Thus, eqns. 5.20 to 5.23 provide convenient general relations for computing enthalpy, internal energy and entropy changes as function of volumetric properties and specific . If a is described by a suitable EOS, these equations may be conveniently integrated to obtain analytical expressions for energy and entropy changes. ------Example 5.1 Derive an expression for enthalpy change of a gas during an assuming using the −= following EOS: P() V b RT (Click here for solution) ------5.5 Residual Property Relations An alternate method of computing energy and entropy changes for real gases involves the definition of residual property. The specific residual property M R is defined as follows: MR = MTP(, ) − Mig (, TP ) ..(5.24)

Where, MTP(, ) is the specific property of a at a given T & P, and Mig (, TP )is the value of the same property if the gas were to behave ideally at the same T & P. Thus for example: VR = VV − ig Using the generalized factor ‘Z’: VR =( ZRT /)( P − RT /) P

VR =( Z − 1) RT / P ..(5.25) The residual properties are usually used for gases only. Using such a property for a (or ) is inconvenient as then it would also include the property change of vapourization (and solidification) which generally are large in magnitude. This detracts from the advantage of working with the residual property as a measure of small corrections to ideal gas behaviour. Thus the use of residual functions is restricted to prediction of real gas behaviour only. To exploit the

4 | Page

concept of residual properties we take advantage of the Gibbs free energy as it can be used as a generating function for other thermodynamic properties. Derivation of Residual functions:

We start from the generic equation:G= H − TS ..(5.2)

And dG= VdP − SdT ..(5.7)

Taking the total differential for the functionG/ RT :

GG1 = − d  dG2 dT ..(5.26) RT RT RT

Substituting eqn. 5.2 and 5.7 in 5.26 we get:

GV H = − d  dP2 dT ..(5.27) RT RT RT

One may write the same equation specifically for an ideal gas, whence:

GVig ig Hig = − d  dP2 dT ..(5.28) RT RT RT

Subtracting eqn. 5.28 from 5.27:

GVRR H R = − d  dP2 dT ..(5.29) RT RT RT

Thus we may write the following further generative relations:

VRR∂(/ G RT ) =  ..(5.30) RT∂ P T

HRR∂(/ G RT ) = −T  ..(5.31) RT ∂T P

SR HG RR And further: = − ..(5.32) RT RT RT

RRR R P GGG G From eqn. 5.29: d = − ; but = 0 ∫0  RT RTPP RT =0 RT P=0

5 | Page

RR GVP Thus: = dP ..(5.33) RT∫0 RT

R G P dP Putting eqn. 5.25 in 5.33: = (Z − 1) ..(5.34) RT ∫0 P

Differentiating eqn. 5.34 w.r.t T in accordance with 5.31 gives:

R HP ∂ Z dP = −T ..(5.35) ∫0  RT∂ TP P

Finally using eqns 5.32, 5.34, and 5.35:

R SPP∂ Z dP dP =−TZ −−( 1) ..(5.36) ∫∫00 R∂ TPP P

The last two equations may be expressed in alternative forms in terms of reduced and :

R HZPr ∂ dP = −T 2 r ..(5.37) r ∫0  RTC ∂TPrr Pr

R SZPPrr∂ dP dP =−TZrr −−( 1) ..(5.38) r ∫∫00 R∂ TPrr Pr Pr

------Example 5.2 Derive an expression for enthalpy change of a gas during an isothermal process assuming using the following EOS: Z = 1 + APr / Tr (Click here for solution) ------5.6 Residual Property Calculation from EOS

From Virial EOS:

Using 5.35 and 2.12 one obtains:

HR P B dB = − ..(5.39) RT R T dT

6 | Page

Next, on substituting eqns. 2.13 – 2.15 in 5.37 the following relations result:

H R 1.097 0.894  = −+ω − Pr 0.083 1.6 0.139 4.2  ..(5.40) RTCrT T r 

Similarly using 5.36 and 2.12 we get:

SR P dB = −  ..(5.40) R R dT

Finally employing eqns. 2.13 – 2.15 in 5.38:

S R 0.675  0.722  =−+ω Pr 1.6 4.2  ..(5.41) RTrr  T 

From Cubic EOS:

One may use the following form of the cubic EOS presented in chapter 2:

RT a P = − ..(2.21) V− b V22 ++ ubV wb

Or equivalently: ZZZ32+α + βγ +=0 ..(2.23)

While the eqns. 5.35 and 5.36 are useful for explicit EOS, they are unsuitable for cubic EOS which are pressure explicit. For the latter type of EOS one may show that the appropriate equations for residual enthalpy and entropy are (see S.I. Sandler, Chemical, Biochemical and Engineering Thermodynamics, ch. 6, 4th Edition, Wiley India, 2006):

R HP1 V ∂ =Z −+1 T− P dV ..(5.42) ∫V =∞  RT RT∂ T V

R S 1 V ∂PR =+−ln Z dV ..(5.43) ∫V =∞  R R∂ TVV

The above relations may be applied to the various cubic EOSs to obtain the necessary residual property relations. The final results are shown below. From RK-EOS:

7 | Page

HR 3a ZB+ =(Z −− 1) ln  ..(5.44) RT 2bRT Z

SR a ZB+ =ln(ZB −− ) ln  ..(5.47) R 2bRT Z

0.42748RTT22α ( ) 0.08664RT where,; a = C RK r b = C PPCC −1/2 α RK()TT r= r

For SRK-EOS:

∂a R Ta− H 1 ∂T ZB+ =(Z −+ 1) − ln  (5.48) RT RT bRT Z

SR 1 ∂+a ZB =ln(ZB −− ) ln  ..(5.49) R bRT∂ T Z

∂aa =− (0.48+− 1.574ωω 0.1762 ) ..(5.51) ∂  T αSRK()T r TT C

0.42748RTT22α ( ) 0.08664RT where,; a = C SRK r b = C PPCC 2 1/2 2 and,αSRK ( T r )=++− [1 (0.48 1.574ωω 0.176 )(1 −Tr )]

For PR-EOS:

∂a Ta− HR ∂T ZB++(1 2 ) =(Z −+ 1) ln  ..(5.52) RT 2 2bRT Z+− B(1 2 )

SR 1 ∂a ZB ++(1 2 ) =ln(ZB −− ) ln  ..(5.53) RT2 2bRT∂  Z+− B(1 2 )

∂aa =− (0.37464+− 1.5422ωω 0.629922 ) ..(5.54) ∂  T α PR()T r TT C

8 | Page

0.45724RTT22α ( ) 0.07779RT where,; a = C PR r b = C PPCC and,α ( T )=++− [1 (0.37464 1.5422ωω 0.629922 )(1 −T 1/2 )] 2 PR r r As part of computing the residual enthalpy or entropy (and internal energy) at any set of temperature and pressure, the or correspondingly Zvap needs to be first computed employing the usual algorithm for solving cubic EOS described in section 2.3.3. ------

Example 5.3 Derive expressions for HR, SR from RK-EOS. (Click here for solution) ------

5.7 Generalized Correlations for computing dH and dS for a real gas:

The approach based on the use of the compressibility factors can be applied to the present instance to evolve generalized correlations for computing enthalpy and entropy changes for gases. We start with the Pitzer-type expression for the :

01 ZZ= +ω Z ..(2.25)

Differentiating with respect to the reduced temperature we get:

∂∂ZZ01   ∂ Z  =  +ω   ..(5.55) ∂∂TTrr ∂ T r PPrr    P r

Thus using eqn. 5.54 we may recast eqns. 5.37 and 5.38 as follows:

PP HZR rr∂∂01  ZdP =−−TT22ω r ..(5.56) rr∫∫∂∂ RTc00 T rP TPrr r Pr

P P SZR rr∂∂01 dP Z dP =− +−01 rr −ω +  And: ∫∫TZr  1 Z ..(5.57) R00∂∂ Tr  PTP rrr  PPrr  

Both the above equations may be rewritten individually as follows:

9 | Page

0 H R (H R ) (H R )1 = + ω ..(5.58) RTc RTc RTc

0 R P (H ) r ∂Z 0 = − 2 Where, Tr ∫  RTcr0 ∂T Pr ..(5.59a)

And:

P ()HZR 11r ∂ dP = − 2 r Tr ∫  RTc 0 ∂TPrr Pr ..(5.59b)

Similarly:

0 R SSRR(S ) ()1 = +ω ..(5.60) RR R

0 R (S ) Pr ∂Z 0 dP =− +−0 r ..(5.61a) ∫ TZr  1 RTP0 ∂ rr Pr

1 R P (S ) r ∂Z1 dP =−+∫ Z1 r R0 ∂ TPrr Pr ..(5.61b)

0 R (H ) ()H R 1 The term in eqn. 5.59a constitutes the first order enthalpy departure, and (in eqn. RTc RTc

5.59b) the second order (with respect to simple fluids) enthalpy departure at specified Tr and Pr. The same is true for the corresponding entropic terms provided by eqns. 5.61a and 5.61b. The evaluation of the integrals in eqns. 5.59 to 5.62 may be carried out assuming an EOS. The most widely used approach is that of Lee and Kesler who employed a modified form of the BWR EOS (eqn. 2.17) to extend their generalized correlation to residual property estimation.

RR01 Figs. 5.1 and 5.2 respectively provide values of ()/,and()/H RTCC H RT respectively.

Similar plots for the entropy terms ()/,and()/SRRR01 SRare also available; however, here the

R plot of the entire entropy term ()−S as function of Tr and Pr is shown in fig. 5.3. Table of 10 | Page

values of all the above (eqns. 5.59 to 5.62) are also available as functions of Tr and Pr at discrete intervals (see for example: J.M. Smith, H.C. Van Ness and M.M. Abbott, Introduction to Chemical Engineering Thermodynamics, 6th ed., McGraw-Hill, 2001.

0 Fig. 5.1 Correlation of − HRC RT drawn from tables of Lee-Kesler (Source: AIChE J., pp. 510-527, 1975)

11 | Page

1 Fig. 5.2 Correlation of − HRC RT drawn from tables of Lee-Kesler (Source: AIChE J., pp. 510-527, 1975)

12 | Page

Figure 5.3 Generalized entropy departure functions using corresponding states. [Source: O.A. Hougen, K.M. Watson, and R.A. Ragatz, Chemical Process Principles Charts, 2nd ed., John Wiley & Sons, New York, 1960]

5.8 Computation of ΔH and ΔS for a Gas using Generalized Departure Functions The residual function equations presented in the last section are particularly useful for estimating finite changes in enthalpy and entropy for real gases undergoing change in either closed or open system processes. We consider that a pure fluid changes state from (TP11 , ) to ( TP 2 , 2 ); shown schematically in fig. 5.4.

13 | Page

Fig. 5.4 Schematic of a General on P – T co-ordinates

Since the departure functions HSRR and capture deviations from ideal gas behaviour at the same temperature as the real gas, one can conceive of the pathway between states ‘1’ and ‘2’ to be decomposed into following steps (see fig 5.5):

a) Real gas state at (,TP11 ) to ideal gas state (ig) at (,TP11 )

b) Ideal gas state at (,TP11 ) to ideal gas state at (,TP22 )

c) Ideal gas state at (,TP22 )to real gas state at (,TP22 )

5.5 Pathway for calculating ΔH and ΔS for Real Gases

ig R • For step ‘a’ the change of enthalpy is given by: HH11−=− H 1 ig ig ig • For step ‘b’ the change of enthalpy is given by: HH21−=∆ H

ig R • For step ‘c’ the change of enthalpy is given by: HH22−= H 2 14 | Page

Therefore, the overall change of enthalpy is given by:

T ∆ = − =R − R +∆ ig2 ig H H21 H H 2 H 1 H∫ Cp dT T1

T −=RR − +2 ig Using eqn. 3.8: H21 H H 2 H 1∫ Cp dT ..(5.63) T1 The same considerations apply for computing the change of entropy between the two states:

R R ig ∆SS =21 − S = S 2 − S 1 +∆ S

T −=RR − +2 ig − Using eqn. 4.21: S2 S 1 S 2 S 1∫ Cp dT/ T R ln( P21 / P ) ..(5.64) T1 Generalized residual property relations may be used for calculation of change in internal energy i.e. UU21− for a process in the following manner:

U2−= U 1( H 2 − PV 2 2 )( − H 1 − PV 11 )

Or: U2−= U 1( H 2 − H 1 )( − PV 2 2 − PV 11 ) ..(5.65)

The term ()HH21− can be calculated using eqn. 5.63, while the term ()PV2 2− PV 11 may be computed after obtainingVV12and applying the generalized compressibility factor approach. One may, however, also use the generalized residual property charts for internal energy for the same purpose (fig. 5.6). ------Example 5.5 Estimate the final temperature and the work required when 1 mol of n-butane is compressed isentropically in a steady-flow process from 1 bar and 50oC to 7.8 bar. (Click here for solution) ------

15 | Page

Figure 5.6 Generalized internal energy departure functions using corresponding states [Source: O.A. Hougen, K.M. Watson, and R.A. Ragatz (1960), Chemical Process Principles Charts, 2nd ed., John Wiley & Sons, New York]

5.9 Extension to Gas Mixtures The generalized equations developed for ∆∆H and S in the last section may be extended to compute corresponding changes for a real gaseous mixture. The method used is the same as that

16 | Page

developed in section 2.4 through the definition of pseudo-critical mixture properties using linear mixing rules: = = ωω= TCm,,∑ yT i Ci PCm,,∑ yP i Ci m∑ y ii ..(2.32) i i i Using the above equations pseudo- are computed:

Trm,,= TT/ Cm and Prm,,= PP/ Cm Further calculations of changes in internal energy, enthalpy, and entropy follow the same principles developed in the last section. ------Example 5.6 Calculate the changes in enthalpy and entropy per when a mixture of 70 mole % ethylene (1) and 30 mole% propylene (2) at 323K and 10 bar is taken to 60 bar and 600 K using the generalized compressibility factor approach.

ig −−3 62 −93 Cp1 =+−+4.196 154.565xT 10 81.076 xT 10 16.813 xT 10 ;

Cig =+−+3.305 235.821xT 10−−3 117.58 xT 1062 22.673 xT 10 −93 p2 (Click here for solution) ------5.10 Relations for ΔH and ΔS for Liquids

One starts with the generic equations for dH and dS developed in section 5.4.

∂V dH= CP dT +− V T dP ..(5.20) ∂P T

dT∂ V dS= CP − dP ..(5.22) TT∂ P

As discussed in section 2.3, for liquids it is often simpler to use volume expansivity and isothermal compressibility parameters for computing thermodynamic properties of interest. Thus from eqn. 2.3: 1 ∂V Volume Expansivity , β ≡  VT∂ P Using eqn. 2.3 in 5.20 and 5.22 the following relations obtain: 17 | Page

dH= CP dT +−[1 β T] dP ..(5.66)

dT dS= C − β VdP ..(5.67) P T

The above equations may also be used to compute the properties of a compressed liquid state. Since the volumetric properties of liquids are very weakly dependent of pressure one can often use the saturated liquid phase properties as reference points and integrate the eqns. 5.66 to 5.67 (at constant temperature) to obtain enthalpy and entropy respectively. The relevant equations are:

P H=+− Hsat V(1β T ) dP ..(5.68) ii∫ sat i i Pi

P S= Ssat − β V dP ..(5.69) i i ∫ sat ii Pi

sat In the last two equations, the molar volume Vi may be set equal to Vi ( liq .) , and the volume expansivity approximated to that at the saturated liquid point at the given temperature.

5.11 Applications to real fluid processes in process plant equipments In a typical process plant one encounters a variety of flow devices such pumps, , turbines, nozzles, diffusers, etc. Such devices are not subject to by design as are heat exchangers, condensers, evaporators, reactors, etc. However, the flow devices typically are subject to mechanical irreversibility owing to existence of dissipative such as fluid viscosity and mechanical friction, which results in reduction of their efficiency. In addition such devices may be subject to thermal irreversibility as their operation may not be truly adiabatic. Therefore, it is necessary to compute the efficiency of such devices in relation to a perfectly reversible (isentropic) process between their inlet and outlet. The performance of a flow device is expressed in terms of isentropic efficiency in which the actual performance of the device is compared with that of an isentropic device for the same inlet conditions and exit pressure. For example, the isentropic efficiency ηT of a turbine (which essentially converts fluid enthalpy to shaft work, fig. 5.7) is defined as:

Power output of the actual turbine HH− = ie η=T s ..(5.69) out put of the turbine,if it were isentropic Hie− H

18 | Page

Fig.5.7 Schematic of a Turbine

Where, Hi = enthalpy of the fluid at the inlet of the turbine, He = enthalpy of the fluid at the exit

s of the actual turbine, He = enthalpy of the fluid at the exit of the turbine, if it were isentropic.

Similarly, the isentropic efficiency ηc of a (fig. 5.8) or a pump η p (which convert applied shaft work to fluid enthalpy) is given by:

s s (Ws ) HHei− ηηcP()or = = ..(5.70) (Ws) HH ei−

Where, Hi = enthalpy of the fluid at the inlet to the compressor (or pump), He = enthalpy of the

s fluid at the exit of the actual compressor (pump), and He = enthalpy of the fluid at the exit of an isentropic compressor (or pump), and Ws represents the shaft work per mole (or ) of fluid in the two situations.

Fig.5.8 Schematic of a Turbine

19 | Page

The isentropic efficiencyηN of a nozzle, which is used to achieve high fluid velocity at exit (by conversion of enthalpy to ) is given by: (V 2 )/2 Kinetic energy of gas leaving the actual nozzle η =2 = N 2 s ..(5.71) (V2 ) /2 Kinetic energy of gas leaving the nozzle, if it were isentropic

------Example 5.7 A certain gas is compressed adiabatically from 293 K and 135 KPa to 550 KPa. What is the work needed? What is the final T2? Assume ideal gas behavior. Compressor η = 0.8. ig -3 -6 2 For the gas: Cp = 1.65 + 8.9 x 10 T – 2.2 x 10 T ------Assignment - Chapter 5

20 | Page