You are on page 1of 147

Chapter 8 Solutions

Engineering and Chemical Thermodynamics 2e

Milo Koretsky
Wyatt Tenhaeff

School of Chemical, Biological, and Environmental Engineering


Oregon State University

milo.koretsky@oregonstate.edu
8.1 No this will not work. A central idea behind binary phase equilibrium is that when both species
are volatile, there will be a mixture of 1 and 2 in both the vapor and the liquid. There will be a
higher mole fraction of the lighter component, toluene, in the liquid relative to the mole in the
vapor, but there will also be cyclohexane present.

2
8.2 Yes if species b is the lighter component, the phase diagram for the system looks approximately
like the one below. We can see that both xa and ya are smaller at the higher pressure, P2:

Pbsat

P2
P1
xa

Pasat
ya

xa ya

3
8.3 The addition of the second species to the liquid will lower the liquid mole fraction of water.
Thus, the fugacity of water in the liquid will be lower than its fugacity in the vapor. Since T and
P are constant, the fugacity of the steam remains constant. Since the fugacity of water in liquid is
lower than vapor, it will all condense.

4
8.4 The addition of the second species to the liquid will lower the liquid mole fraction of water.
Thus, the fugacity of water in the liquid will be lower than its fugacity in the vapor and the steam
will condense. However, since the container is rigid, the pressure of the vapor will decrease as
the steam condenses, lowering its fugacity. Therefore, some of the steam will condense until the
pressure in the vapor is low enough that the fugacity in the vapor matches the fugacity in the
liquid.

5
8.5 When the third component is added to the liquid, the mole fraction of the water in the liquid
decreases so tis fugacity decreases. Thus, the fugacity of water in the vapor will be greater than
in the liquid and some water will condense, lowering the number of moles in the vapor.

6
8.6

(a) since i <1 (ln i < 0), the unlike interactions are stronger and the like interactions are weaker.

(b)
1.4 1.4
1.2 1.2
1.0 1.0
0.8 0.8
P [bar]
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
0.0 0.2 0.4 0.6 0.8 1.0
xa, ya

Since the pressure with a-b interactions is higher than the ideal solution case where all the
interactions are the same (blue line, Raoult’s law), the like interactions are stronger.

(c) Since it takes more energy to boil the mixture (maximum boiling azeotrope), the unlike
interactions are stronger and the like interactions are weaker.

7
8.7

(a)
(i).
For positive deviations from ideality, both activity coefficients are greater than unity. This
implies that the like interactions of the species are stronger than the unlike interactions, which is
true for a mixture of methane and ethanol. Ethanol con hydrogen bond with itself, which leads
to strong like-like interactions. The H bond interactions are much greater than the dispersion and
induction of ethanol – methane.

(ii).
Both activity coefficients are less than one in this case, so the unlike interaction are stronger than
the like interactions. This behavior is seen in acetone/chloroform mixtures because the partially
positive hydrogen atom in chloroform is attracted to the partially negative oxygen in acetone
leading to a H bond in the unlike species. Neither of the like interactions are nearly this strong.

(b)
Ethane/Methane. Since methane and ethane are similar in structure, the unlike interactions are
very similar to like interactions.

(c)
The cross coefficient, B12, is more negative than the pure species B parameters when the unlike
interactions are stronger than either of the like interactions. This statement is true for acetone
and chloroform (see (a)(ii)).

(d)
The geometric average is an accurate estimate of the cross virial coefficient when dispersion
forces govern the interactions between unlike molecules and the polarizabilities of the two
species are nearly equal. This is true for methane/ethane mixtures.

8
8.8
In System I at the composition where the two g curves touch, the Gibbs energies are equal and
we have vapor-liquid equilibrium. At all other compositions for this T and P, the Gibbs energy
of the liquid phase is lower so the liquid more stable. This behavior is consistent with a
minimum boiling azeotrope (see Figure 8.8(a)), where the temperature of the system is the
temperature of the azeotrope.

In System II at the composition where the two g curves touch, the Gibbs energies are equal and
we have vapor-liquid equilibrium. At all other compositions for this T and P, the Gibbs energy
of the vapor phase is lower so the vapor more stable. This behavior is consistent with a
maximum boiling azeotrope (see Figure 8.8(b)).

9
8.9
(a) False – the increase in entropy upon mixing can allow two species to mix when the unlike
interactions are weaker.

(b) True – the energetic effects outweigh the increase in entropy upon mixing

10
8.10
The purpose of applying salt to the road is to lower the freezing point of water. Therefore, we
want to find the lowest temperature at which liquid water exists. From the diagram, the lowest
temperature is -21.12 ºC. The smallest composition of salt required to lower the freezing point to
this temperature is approximately 23 wt%.

11
8.11

For increasing xa in System 1:


 Pure  until first tie line is reached
 Mixture of  and  along the tie line
 Pure  until the second tie line is reached
 Mixture of  and  along the second tie line
 Pure  for the remaining compositions

For increasing xa in System 2:


 Pure  until the first line is reached
 Pure  and  along the tie line
 Pure  for the remaining compositions

(Along the tie lines, the compositions are determined using the lever rule.)

12
8.12
The species that dissociate into ions will lower the fugacity of liquid water the
greatest. More ions with MgCl2.

Highest boiling point


0.05 M MgCl2
0.05 M NaCl
0.05 M glucose (C6H12O6)
pure water
Lowest boiling point

13
8.13
The species that dissociate into ions will lower the fugacity of water the greatest.
More ions with MgCl2.

Highest freezing point


pure water
0.05 M glucose (C6H12O6)
0.05 M NaCl
0.05 M MgCl2
Lowest freezing point

14
8.16
A Txy phase diagram displays the bubble point and dew point over the entire range of liquid and
vapor mole fractions. By assuming ideal solution and ideal gas behavior of the mixture, we can
use the following relationships

P  x1P1sat  1  x1  P2sat

y1P 1  y1  P
1 
P1sat P2sat

We can calculate the saturation pressures using Antoine’s Equation. The temperature that
satisfies the first equation is the bubble point. The temperature that satisfies the second equation
is the dew point. The following spreadsheet was created with these two expressions (the
temperatures were found using a “Solver” function):

x1 Tbp (K) y1 Tdp (K)


0 341.5 0 341.5
0.1 342.5 0.1 342.9
0.2 343.5 0.2 344.2
0.3 344.6 0.3 345.5
0.4 345.7 0.4 346.8
0.5 346.9 0.5 348.0
0.6 348.1 0.6 349.2
0.7 349.4 0.7 350.3
0.8 350.7 0.8 351.4
0.9 352.0 0.9 352.4
1 353.4 1 353.4

The data were plotted to obtain the following graph:

Txy Diagram for Cyclohexane and n-Hexane

354.0
352.0
Temperature (K)

350.0
348.0
346.0
344.0
342.0
340.0
0.00 0.20 0.40 0.60 0.80 1.00
x1, y1

15
8.17
The vapor mole fractions of n-pentane (1) and n-hexane (2) are

y1  0.33
y2  0.67

We can solve this problem by finding the temperature at which the following relationship is true:

y1 P y2 P
 1
P1sat P2sat

The saturation pressures can be found using Antoine’s Equation. From Appendix A,

Species A B C
n-pentane (1) 9.2131 2477.07 -39.94
n-hexane (2) 9.2164 2697.55 -48.78

Therefore, we must solve the following equation for T:

 0.331 bar  
 0.67 1 bar  1
 2477.07   2697.55 
exp  9.2131   exp  9.2164  
 T  39.94   T  48.78 

We obtain

T  334 K

16
8.18
To calculate the total pressure and the vapor composition we can use Equations 8.8 and 8.9.
These equations require using saturation pressures, which can be calculated with Antoine’s
Equation. From Table A.1.1,

Species A B C
Propylene (1) 9.0825 1807.53 -26.15
Propane (2) 9.1058 1872.46 -25.16
n-butane (3) 9.0580 2154.90 -34.42
Isobutane (4) 8.9179 2032.76 -33.15
n-pentane (5) 9.2131 2477.07 -39.94

Therefore, the saturation pressures at 250 K are

Species P sat bar 


Propylene (1) 2.74
Propane (2) 2.18
n-butane (3) 0.391
Isobutane (4) 0.634
n-pentane (5) 0.076

Equation 8.8:

P  x1P1sat  x2 P2sat  x3 P3sat  x4 P4sat  x5 P5sat


P  0.1 2.74 bar   0.25  2.18 bar   0.2  0.391 bar   0.35  0.634 bar   0.1 0.076 bar 
P  1.13  bar 
Equation 8.9:

x P sat
yi  i i
P

Using this equation, the following table can be created

Species yi
Propylene (1) 0.242
Propane (2) 0.482
n-butane (3) 0.069
Isobutane (4) 0.196
n-pentane (5) 0.0067

17
8.19
Naming Protocol:

Cyclohexane Species 1
Benzene Species 2
Toluene Species 3
n-heptane Species 4

To determine the temperature and vapor compositions, the saturation pressure of each species is
required. The vapor pressures can be calculated Antoine’s equation and data from Appendix A:

 2766.63 
P1sat  exp  9.1325  
 T  50.50 
 2788.51 
P2sat  exp  9.2806  
 T  52.36 
 3096.52 
P3sat  exp  9.3935  
 T  53.67 
 2911.32 
P4sat  exp  9.2535  
 T  56.51 

Expression for pressure assuming ideal gas behavior:

P  x1P1sat  x2 P2sat  x3 P3sat  x4 P4sat

Substitute the saturation pressure expressions and solve for T:

T  361.3 K

At this temperature

P1sat  1.26 bar P2sat  1.29 bar P3sat  0.511 bar P4sat  0.742 bar

To determine the vapor composition, we can use

x i Pisat
yi 
P

Therefore,

y1  0.504 y 2  0.258 y3  0.128 y 4  0.11

18
8.20
Let the subscript “a” designate n-butane and “b” designate isobutane. First, perform mole
balances:

x a , feed F  x a L  y aV
xb, feed F  xb L  ybV

Using the information provided in the problem statement,

V  0.4 F L  0.6 F

xa, feed  0.6xa  0.40 ya


xb, feed  0.6 xb  0.40 yb

Equation 8.9 states:

xa Pasat xb Pbsat
ya  yb 
P P

Substituting these expressions into the mole balances, we obtain

 x P sat   0.4 Pasat 


xa , feed  0.6 xa  0.40  a a   xa  0.6  
 P   P 
 x P sat   0.4 Pasat 
xb , feed  0.6 xb  0.40  b b   xb  0.6  
 P   P 

We also know that xa  xb  1 . Therefore,

xa , feed xb , feed
sat
 1
0.4 P 0.4 Pbsat
0.6  a
0.6 
P P

For the saturation pressures, we can substitute Antoine’s Equation. The Antoine coefficients can
be found in Appendix A.1.

0.5 0.5
 1
 2154.9   2032.76 
0.4 exp  9.058   0.4 exp  8.9179  
 T  34.42   T  33.15 
0.6  0.6 
1 bar 1 bar

Solve for T:

19
T  266.6 K

At this temperature

Pasat  0.8 bar Pbsat  1.23 bar

Therefore,

y a  0.8 x a yb  1.23 xb

We can obtain the following equation from the mole balance on species a:

x a, feed  0.6 x a  0.40.8 x a 


0.5  0.92 x a
 xa  0.543
xb  1  x a  0.457

Now calculate the vapor mole fractions

y a  0.8 x a
y a  0.80.543  0.434
yb  0.566

20
8.21
A mass balance on component i gives

x i, feed F  y iV  x i L (1)

If we assume ideal gas and ideal solution, the equilibrium relation in the flash drum can be
written according to Raoult’s law:

yiP  xiPisat (2)

Substituting Equation 2 into 1 gives

x P sat  P sat 
xi, feed F  i i V  xi L  xi  i V  L 
P  P 
 
Solving for xi:

xi, feed
xi  (3)
Pisat V   L 
  
P F F

An overall balance gives

F V  L (4)

Since the sum of the mole fractions equals 1, we have


xi , feed
1   xi   (5)
Pisat  V   L 
  
P F F

Using Equation 4, we get

xi , feed
1   xi  
Pisat V   V 
   1  
P F  F

This leaves 1 equation for the unknown (V/F). Once V/F is determined, xi is found from
Equation 3. Then yi is found from Equation 2. The solution summary is illustrated below:

21
Antoine coefficients a b c
nC4H10 nC5H12 nC6H14
A 9.058 9.2131 9.2164
B 2154.9 2477.07 2697.55
C -34.42 -39.94 -48.78
Initial conditions
T [K] 290
P 0.6

sat
calc P
Psat 1.8712 0.5002 0.1399

calc xi
xi 0.1988 0.3259 0.4734
calc yi
yi 0.6200 0.2717 0.1104

l/f 0.5224
v/f 0.4776

22
8.22
Since the pressure is 0.5 bar, the vapor can also be assumed ideal. Therefore, the vapor
compositions can be calculated as follows:

x a Pasat xb Pbsat
ya  yb 
P P

To find the saturation pressures, we use Antoine’s Equation and the appropriate values from
Appendix A. We obtain,

Pasat  1.03 bar Pbsat  0.060 bar

Since the solution is ideal, we also have the following relationship:

P  xa Pasat  1  xa  Pbsat
0.5 bar  x a 1.03 bar  1  xa 0.060 bar 

 xa  0.453 xb  0.547

Now we can calculate the gas compositions:

ya 
0.4531.03 bar   0.933
0.5 bar
yb  1  y a  0.067

23
8.23
Since the pressure is 0.333 bar, we will assume ideal gas behavior. The following relationships
hold:

ya P  xa a Pasat
yb P  xb b Pbsat
P  xa a Pasat  1  xa  b Pbsat

First, find the saturation pressures using data in Appendix A.

Pasat  0.484 bar Pbsat  0.0824 bar

Now, create expressions for the activity coefficients using the equations in Table 7.1:

 1816 1 x  2   1816  x  2 


 a  exp  a
  b  exp  a

 8.314  289.15   8.314  289.15 

We have one equation for one unknown:

  1816 1 x 2    1816 x 2 
    a
  0.484 bar   1 xa   exp  
   a   0.0824 bar
  8.314  289.15   
0.333 bar  xa exp 
  8.314  289.15 
     

xa  0.477
We can substitute the liquid mole fraction of n-pentane into the following equations to obtain the
vapor compositions

  1816  1 x  2 
xa  exp  a
  0.484 bar 
   8.314   289.15 
 
ya 
0.333 bar

Therefore,

y a  0.847
yb  1  y a  0.153

24
8.24

(a)
Calculate the activity coefficients from A and B:

 1 
 a  exp  A  3B x b2  4Bx b3 
RT 
 1  3 
 b  exp   A  3B  1 xb   4B 1 xb  
2

 RT  

Calculate the saturation pressures with Antoine’s equation:

Species A B C P sat (Pa)


Isobutane (a) 8.9179 2032.76 -33.15 1.83 10 
Hydrogen
Sulfide (b)
9.1058 1872.46 -25.16 11.6  10 5

We can find the mole fractions using the following equation:

P  xa a Pasat  xb b Pbsat


 1    1  3  sat
P  1 xb  exp   A  3B xb2  4Bxb3  Pasat  xb exp         
2

 RT    RT 
A 3B 1 x b 4B 1 x b  Pb

Substituting values, we obtain

xb  0.47
 xa  0.53

Calculate the activity coefficients:

 a  1.07
 b  1.42

Now, calculate the vapor mole fractions:

x  P sat  0.53 1.07 1.8310  Pa 


5

ya  a a a   0.118
P 8.77 10 5  Pa 
yb  1  y a  1  0.118  0.882

25
(b)
We can use the following equilibrium expressions to calculate equilibrium concentrations:

ya a P  xa a asat Pasat yb b P  xb b bsat Pbsat

These equations can be rearranged to yield

xa aasat Pasat xb bbsat Pbsat


P P
yaa ybb

Use the fugacity coefficient expressions from Table 7.1:

 aa  P   ab  P 
 a  exp ba    b  exp bb   
 RT  RT   RT  RT 

The a and b parameters can be calculated with the van der Waals EOS:

aa  1.33 ba  1.16  10 4
ab  0.454 bb  4.34  10 5

Substitute these values into the expressions for the fugacity coefficient:

 a  0.84  asat  0.964


b  0.943  bsat  0.926

Substitute numerical values and activity coefficient expressions into the equations for pressure:

 
  A  3B x b2  4Bx b3 1.83 10 5 0.964 
1
1 x b  exp
8.314 277.65  
8.77 10 5 
0.84 y a
 
x b exp
1
8.314 277.65
 2 3



 A  3B1 x b   4B1 x b  11.6 10 5 0.926
8.77 10 5 
0.9431 y a 

We have two equations for two unknowns. Solve simultaneously:

x b  0.47  x a  1 0.47  0.53


y a  0.14  y b  1 0.14  0.86

26
8.25
The following conditions hold at the azeotrope:

 1P1sat  P   1P1sat

We can find the partial pressure of ethanol at 60 ºC using Antoine’ equation:

B
ln P1sat  A 
T C

The A, B, and C parameters can be found in Appendix A. Using these values, we calculate at 60
ºC:

P1sat  0.468 bar P2sat  0.56 bar

Therefore,

P 0.64 bar P 0.64 bar


1    1.37 and  2    1.14
P1sat 0.468 bar P2sat 0.56 bar

To determine A, we can use the appropriate equation from Table 7.1.

For species 1
  J 
A  0.6    8.314  mol  K    333.15 K  ln 1.37 
2

 
 J 
A  2422 
 mol 

For species 2
  J 
A0.42   8.314  333.15 K  ln 1.14
  mol  K  
 J 
A  2311 
 mol 

The best value is obtained by averaging these two:

 J 
A  2366 
 mol 

27
(b)
We can use Equation 8.16 to solve for the vapor mole fractions, but first find the activity
coefficients at the given liquid composition:

  2366   0.2 2    2366   0.8 2 


1  exp    1.03  2  exp    1.73
 8.314  333.15   8.314 333.15 

Equation 8.16:

x11P1sat 0.8 1.03  0.468


y1  
x11P1sat  x2 2 P2sat 0.8 1.03  0.468  0.2 1.73  0.56 

y1  0.67
y 2  1  y1  0.33

28
8.26

(a)
The curve that goes through the origin should be labeled with  2 and the other curve belongs to
 1. This is the case because for a Lewis-Randall reference state,

As x1  1,  1  1  ln  1  0

(b)
The like forces are stronger because the activity coefficients are greater than one. The fugacity is
greater than the saturation pressure for a mixture; therefore, we can infer that more molecules are
volatilizing due to the weaker unlike interactions.

(c)
The total system pressure is calculated as follows:

P  x1 1 P1sat  x 2 2 P2sat

From the steam tables,

P2sat  1.013 bar

The diagram in the problem statement provides

ln  1  1.05   1  2.86
ln  2  0.15   2  1.16

Substitute and calculate:

P  0.2  2.86 1.12 bar   0.8 1.16 1.013 bar 


P  1.58  bar 

(d)
We can assume the vapor phase is ideal since the pressure is relatively low. For equilibrium to
exist, the fugacities of the liquid and vapor phases must be equal.

y1 P  x1 1 P1sat

Therefore,

x1 1 P1sat 0.2  2.86 1.12 bar 


y1  
P 1.58 bar

29
y1  0.41

(e)
From the graph, we can find the activity coefficient of species 1 at infinite dilution.

ln  1  2.6   1  13.5

Now calculate the Henry’s law constant with

[1
 1 
P1sat
[ 1  13.51.12 bar   15.1 bar 

(f)
This system does exhibit an azeotrope. Because the pressure calculated in Part (c) is greater than
either of the pure species saturation pressures, the pressure must go through a maximum, where
an azeotrope will form.

30
8.27
An iterative technique will be used to solve this problem. We are assuming ideal solution
behavior. We calculate the pressure using

y1
ˆ1P  x11sat P1sat
ˆ 2 P  1 y1 
y 2 ˆ 2 P  x 2 2sat P2sat

We can combine the equations to obtain:

x11sat P1sat x 2 2sat P2sat


P 

ˆ1 ˆ2

We can calculate the fugacity coefficients using the equation from Table 7.1:

1sat  0.755
 2sat  0.859

For the first round of calculations, we will assume the fugacity coefficients are equal to one.
Substituting values from the problem statement, we get

P  18.4 bar

Now we can calculate the vapor mole fractions:

x11sat P1sat
y1  y 2  1  y1

ˆ1P

We obtain:

y1  0.722 y 2  0.278

Using the mole fractions and pressure obtained above calculate the fugacity coefficients from
Table 7.1:

 P  v  bmix  
ln ˆ1v    ln  
b1 
2 y1a1  y2 a1a2 
 RT   v  bmix  RTv
 P  v  bmix  
ln ˆ2v   ln  
b2 

2 y2 a2  y1 a1a2 
 RT   v  bmix  RTv

Note: a mix  y12 a1  2 y1 y 2 a12  y 22 a 2


bmix  y1b1  y 2 b2

31
For the mole fractions shown above and critical data in Appendix A,

amix  2.26
bmix  0.000136

Therefore,


ˆ1v  0.876 
ˆ 2v  0.792

Now, repeat the process above starting with the new fugacity coefficients. We obtain

P  21.6 bar
y1  0.702 y2  0.279
ˆ  0.854
1
v
ˆ2v  0.664

Iterating a third time, we get

P  22.3bar
y1  0.697 y 2  0.298

ˆ1v  0.849 
ˆ 2v  0.745

A fourth iteration provides

P  22.5 bar
y1  0.695 y 2  0.305

ˆ1v  0.848 
ˆ 2v  0.743

The fifth iteration results in

P  22.5 bar
y1  0.695 y 2  0.305

Since the pressure for the fifth iteration is essentially equal to the pressure obtained in the fourth
repetition, we have solved the problem. The final answer to this problem is

P  22.5 bar
y1  0.695
y 2  0.305

32
8.28
As in Example 8.6, an iterative technique is required to solve this problem.

Step 1.
Determine all of the constants for the problem. Calculating the pure species a and b parameters
(a1, b1, a2, and b2) for the van der Waals equation is described in Chapter 4. Critical values are
provided in Table A.1.1. To find the a and b parameters for the mixture, use

a mix  y12 a1  2 y1 y 2 a12  y 22 a 2


bmix  y1b1  y 2 b2

The activity coefficients will also remain constant. Calculate them with the following equations:

  A  3B  2 4 B 3    A  3B  2 4 B 3 
 1  exp  x2  x2   2  exp  x1  x1 
 RT RT   RT RT 

The saturation pressures can be calculated using Antoine’s Equation.

Step 2.
Set the fugacity coefficients equal to one for the initial calculation.

Step 3.
Calculate the pressure using

x1 1 P1sat x 2 2 P2sat


P 
ˆ1 ˆ 2

Step 4.
Evaluate the vapor mole fractions with the following equations:

x1 1 P1sat
y1  y 2  1  y1
̂1 P

Step 5.
Now that the vapor mole fractions are known, calculate the fugacity coefficients with the
equation from Table 7.1:

ln ̂1v   ln 

P vb
mix    b1


2 y1a1  y2 a1a2 
 RT  v  b mix
RTv

33
 v
ln ̂2   ln 

P vb
mix    b2


2 y2 a2  y1 a1a2 
 RT  v  b mix
RTv

Note: You must solve for v with van der Waals EOS (use the pressure obtained in Step
3).

Step 6.
Start a new iteration at Step 3, using the new values for fugacity coefficients. Repeat steps 3
through 6 until the change in pressure from one iteration to another is smaller than your
acceptable limit.

34
8.29
(a) We can examine the Margules parameter A and we find that A > 0. This indicates that like
interactions are stronger than unlike interactions.

(b)
P  xa a Pasat  xb b Pbsat

RT ln  a   A  3B  xb2  4 Bxb3  3174.4 J/mol


RT ln  b   A  3B  xa2  4 Bxa3  334.4 J/mol

 a  3.60
 b  1.14

From the Antoine equation:


 3166.38 
Pasat  exp10.9237   0.0270 bar 
 298.15  80.15 

 3816.44 
Pbsat  exp11.6834   0.0312 bar 
 298.15  46.13 

Solving for pressure:


P  0.048 bar

(c)
Solving for mole fraction
xa a Pasat
ya   0.41
P

35
8.30

Determine the temperature and the vapor phase of mole fraction a:

J
0.6 2.74 7600 mol

74.5 kPa exp

Low P ideal gas

P P

P P

A A
P P P exp P exp P

74.5 kPa 0.6exp 2.74 0.4 0.4exp 2.74 0.6

→ steam tables

→ Antoine equation

2788.51
ln P 9.2806
52.36

Guess and check:


TC P kPa P kPa P(kPa)

50C 12.35 36 50.3

70C 31.19 73 108

*60C 19.94 52 74.5

P
0.25

36
8.31
(a)
Given:
x1  0.471
x2  1  x1  0.529
P  0.073 bar

From the Antoine Equation at 20 oC:

P1sat  0.10 bar


P2sat  0.29 bar

Setting fugacities equal:

y1 P  x1 1 P1sat
y2 P  x2 2 P2sat
P  x1 1 P1sat  x2 2 P2sat

Applying the two-suffix Margules Equation

 A 2
 1  exp  x2 
 RT 
 A 2
 2  exp  x1 
 RT 

So

 A 2  sat  A 2  sat
P  x1 exp  x2  P1  x2 exp  x1  P2
 RT   RT 
Solving for A:

 J 
A  1220 
 mol 

(b)
We can examine the Margules parameter A and we find that A > 0. This indicates that like
interactions are stronger than unlike interactions.

37
(c)
 A 
 1   2  exp    1.68
 RT 

[ 1   1 P1sat  0.172 bar

[ 2   2 P2sat  0.048 bar

(d)
x1 1 P1sat
y1   0.77
P

(e)
Lever rule
nl y z 0.77  0.7
 1 1  0.23
nT
y1  x1 0.77  0.471

nl  2.3 mol

38
8.32
To calculate the pressure, we can use the following:

To calculate the mole fraction of vapor we use:

Start by calculating & , using the three-suffix Margules model

RT ln  a   A  3B xb2  4 Bxb3
RT ln  b   A  3B xa2  4 Bxa3

Next we need to use the Antoine equation to calculate Psat for each of the species in our solution.

 3166.38 
Pasat  exp10.9237   0.0270 bar 
 298.15  80.15 

 3816.44 
Pbsat  exp11.6834   0.0312 bar 
 298.15  46.13 

This can be determined by plotting xa and ya vs P
0.06

0.05

0.04
Pressure [bar]

0.03

0.02

0.01

0.00
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
xa, ya

Where xa is the blue line and ya is the red line. We can see that it does form an azeotrope. Next
we can determine the pressure at which that azeotrope occurs via solver or other analysis.

P = 0.0477 bar

39
8.33
At the azeotrope

x1  y1 x2  y 2

Therefore, the expressions equating liquid and vapor fugacities simplify to

P   1 P1sat P   2 P2sat
(Note: We are also assuming that the vapor behaves ideally, which is reasonable since the
pressure is 122.3 torr.)

We can calculate the saturation pressures at 25 ºC using Antoine’s equation data in Appendix A.

 3803.98 
P1sat  exp 12.2917   0.078 bar  58.5 torr 
 298.15  41.68 
 2788.51 
P2sat  exp 9.2806   0.127 bar  95.2 torr 
 298.15  52.36 

For the van Laar equation

 A  Bx 2 
2  B  Ax1 
2
 1  exp      2  exp    
 RT
  Ax1  Bx2  

 RT
  Ax1  Bx2  

Substitute these expressions into the above equations for pressure. We have

 P  2  P  2
ln   A  Bx 2 
 ln   B  Ax1 

 P sat  RT  Ax1  Bx 2   P sat  RT  Ax1  Bx 2 
 1   2 

Inserting numerical values, we have two equations for two unknowns:

 B  0.72 
2
 122.3  A
ln    
 58.5  8.314  298.15  A  0.28  B  0.72 
 A  0.28 
2
 122.3  B
ln    
 95.2  8.314  298.15  A  0.28  B  0.72 

Solve the equations simultaneously:

40
 J 
A  6416.0 
 mol 
 J 
B  2856.6  
 mol 

Now, to calculate the pressure and liquid composition when the vapor mole fraction of ethanol is
0.75, we can use the following equations

y1 P  x1 1 P1sat y 2 P  x2 2 P2sat

Therefore,

y1 P y1 x1 1 P1sat
 
y 2 P y 2 x2 2 P sat
2
y1 P2sat x
 1 1
y 2 P1sat x2 2

Substitute the activity coefficient expressions for the van Laar equation:

  2
 Bx2   A  B 1 x1   
2

x1 exp 
A
  x1 exp   
 RT  Ax1  Bx2    RT  Ax1  B 1 x1   
y1P2sat  
 
   
y2 P1sat B  Ax1   
2 2

  B Ax
x2 exp   1 x1  exp  1
 
 RT  Ax1  Bx2   RT  Ax1  B 1 x1   
 

Insert numerical values and solve for x1:

x1  0.935
 x 2  0.065

Now, we can calculate the pressure


  
2856.6  0.065
2

 6416
 0.935  58.5 torr  exp    
        
x11P1sat 8.314 298.15 6416 0.935  2856.6 0.065
P 
y1 0.75

P  73.1  torr 

41
8.34

(a)
The molar volume can be found by manipulating the expression given in the problem statement:

 PAy1 y 2  B 
RT
v
P
8.314  298.15  90 10 5 8.314 298.15  2 1014 1 / 3 2 / 3  81015 
v
90 10 5
  
  
    
 m3 
v  3.55  10  4  
 mol 

We calculate the total volume as follows

V  nT v  15 mol 3.55 104 m 3 / mol  5.3310 3 m 3 

The molar volume of species 2 is calculated by letting y 2  1.

8.314  298.15 
v2 
90 10 5  90 10  8.314  298.152 10   0 1  810
5 14 15 

 m3 
v 2  4.54  10  4  
 mol 

The partial molar volume of species 2 is calculated by evaluating the appropriate derivative:

 V    n1  n2 RT  nn 
V2       P  A 1 2  Bn1  n2  
 n2 T , P, n1 n2  P  n1  n2 

 P  Ay12  B
RT
V2 
P
 m3 
V2  4.04  10  4  
 mol 

(b)
To find the pure species fugacity coefficient, first find an expression for the pure species
fugacity:

P
 f 2v 
 v2 dP  RT ln  Plow 
Plow

42
P
 RT   fv 
  P  PB dP  RT ln  2 
  Plow 
Plow

 P  B 2  f2v 
RT ln   
 Plow  2
 P  Plow 
2
 RT ln  
 Plow 

If we let Plow go to zero, cancel the remaining Plow’s, and simplify, we obtain

 B 2
f2v  P exp 
 2RT 
P

 B 2
 2v  exp  P  1.38
 2RT 

To find the fugacity coefficient of species 2 in the mixture, we use the following procedure:

P  fˆ v 
 V2 dP  RT ln  y 2 Plow 
2

Plow  
P
 RT   fˆ v 
  P  1 
 P Ay 2
 B dP  RT ln  2 
Plow  y2 Plow 
 P  1 2  fˆ v 
RT ln     P  Plow   Ay1  B  RT ln  2 
2 2

 low  2
P  y2 Plow 

If we let Plow go to zero, cancel the remaining Plow’s, and simplify, we obtain

 2 
ˆf v  y P exp  P  Ay 2  B 
2 2 1
 2RT 
 P2 
 ˆ2v  exp 
 2RT
 Ay12  B  1.26

(c)
Plot the activity coefficient of species 2 versus its liquid mole fraction:

43
Henry's
Plot of Activity Coefficient  2

1.2
1
0.8
Henry's

0.6
2

0.4
0.2
0
0 0.2 0.4 0.6 0.8 1 1.2
x2

The activity coefficient based on Henry’s Law is less than one. Thus, the fugacity is less than
the fugacity based solely on 1-2 interactions. Consequently, the like interactions must be
stronger.

(d)
For phase equilibrium of species 2 to exist, the following must be true:

y2ˆ2v P  x2 2Henry ' s[ 2

However, we can’t calculate the activity coefficient until we know the liquid mole fraction.
Therefore, we need to make an expression for the activity coefficient containing the mole
fraction:

y2ˆ2v P y2ˆ2v P
 x2  Henry's 
2
 2 
H 2 exp 7 1 1 x  2  H
 2 
We can find the mole fraction with a solver function:

x2  0.013

(e)
Initially assume the fugacity coefficient is one at the saturation pressure and the Poynting
correction is negligible. Use the following equation:

P2sat
 2Henry ' s 
[2
P   7000 bar  exp 7 1 0 2   6.38  bar 
2
sat

44
The fugacity coefficient at saturation can be found according the the result of part (b):

 B 2
2sat  exp  P  1.00
 2RT 

so the assumption is valid.

45
8.35
At the azeotrope

xa  y a xb  y b

Therefore, the expressions equating liquid and vapor fugacities simplify to

P   a Pasat
P   b Pbsat
(Note: We are also assuming that the vapor behaves ideally.)

Since the liquid-phase nonideality is represented by the two-suffix Margules equation, the
following expressions are used to calculate the activity coefficients:

 A 2  A 2
 a  exp  xb   exp 
 RT   RT
1 xa  

 A 2
 b  exp 
 RT 
xa

Substitute these expressions into the pressure equations and equate the pressures. Solve for xa:

 A 2  A 2
Pasat exp 
 RT
1 xa    Pbsat exp 
  RT 
xa

 2900   2900 
 68.8 kPa  exp 1 xa     46.5 kPa  exp 
2
xa2 
 8.314  328.15   8.314 328.15 
xa  0.68

Now, calculate the activity coefficient of methanol.

 2900  0.32  2 
 a  exp    1.11
 8.314  328.15 

Use this value to calculate the pressure.

P 1.11 68.8  kPa   76.4  kPa 

46
Thus, an azeotrope forms at a pressure of 76.4 kPa and a temperature of 55 ºC. The composition
is 68% methanol and 32% ethyl acetate. This is a minimum-boiling azeotrope (positive
deviations from Raoult’s Law) because the azeotropic pressure is greater than the two saturation
pressures.

47
8.36
J
RTln 8.314 70 273.15 ln99.7 13,130 mol

RT 4.60

RTln 1

1 →

exp 1 1
RT

0.011, 0.75, 1

The value 0.75 makes physical sense.

48
8.37

P  20.57bar
System Properties:
T  301o C  574.15o K

Antoine
A  9.3935 TC  591.7o K
Toluene Data:
B  3096.52 PC  41.14bar
C  53.67   0.257

P
PR   0.5
fˆ  f  1 P
1
v
1
v
PC
1  general correlation 574.15
TR   0.77
591.7

From Appendix C log  ( o )   0.086 and log  (1)   0.023

log 1  log  ( o )   log  (1)  0.092

1  0.810

Vapor is pure, liquid is a binary mixture

f1  1 P  f1v  fˆ1l  x1 11sat P1sat

f1v 16.6
x1 1   sat
1 P1
sat sat
1 31.3

To find 1sat we need to linearly interpolate

31.3 574.15
PRsat   0.761 , TR   0.77
41.14 591.7

PR 0.7 0.8 0.76


log  (o) -0.126 -0.148 -0.139
log  (1) -0.034 -0.042 -0.039

log 1sat  0.149

49
1sat  0.709

x1 1  0.75

From the graph, and more linear interpolation

x1 1 x1 1
0.7 1.06 0.742
0.75 1.04 0.78

x1  0.71

50
8.38
ln ln → 2 suffix Margules → ln 0.75

J
RT ln 1953 mol

0.4 low , , , , →1

80kPa
60kPa

(a)
exp 1 1 exp
RT RT

80.3 kPa 0.45

10 mol

4 mol

10 4

10 4
3.90 mol

(b)
At azeotrope.

8
exp 1
6 RT

RT
ln 1.33 1 2

0.38 1 2

1 0.38
2 0.69
2

51
exp 1 85.9kPa
RT

(c)
Since A>0, the like interactions are stronger

52
8.39
(a)
If this mixture is in equilibrium with vapor under these conditions, what is the vapor phase mole
fraction of a?

J
1 mol A 4010 mol

J
3 mol Table 7.1 B 2501 mol

4 mol RTln A 1703

T 25

P 0.50 bar 50 kPa ln 0.687

0.25 1.99

0.75

P 0.25 1.99 75 kPa


0.745
P 50

(b)
Determine saturation pressure of b.

P P P

A
RTln B 303
A B

ln 0.122
1.13

50 0.25 1.99 75
P 15 kPa
0.75 1.13

53
8.40

(a)
Plot shows Pxy phase diagram for a binary mixture of species 1 and 2 at 300K.

Vapor, y1=0.67

(b)
P ~ 50 kPa
yA~0.77
xA~0.47

54
8.41
(a)
J
7850 mol 0.6

J
3410 mol 0.4

A J
RTln B 2050 mol
A B

2050
ln 0.679
8.314363

From steam tables: 70.1 kPa

1.97

0.4 1.97 70.1 kPa 55.3 kPa

(b)
55 kPa

(c)
[ [

→0

J
RTln 3410 mol

ln 1.13

3.09

[ 3.09 70.1kPa 217 kPa

55
8.42

(a)
1 bar

(b)
(i) liquid
(ii) x1 = 0.2
(iii) 1 mol

(c)
z1 = 0.6

(i) 2 phases, vapor and liquid


(ii) x1 = 0.4, y1 = 0.74
(iii) apply the lever rule

nl 0.74  0.6
  0.7
nv 0.6  0.4
nl  0.82 mol
n v  1.18 mol

(d)

y1 P  x1 1 P1sat

x1 P1sat
1   0.74
y1 P

56
RT ln  1 J
A 2
 2, 040
x2 mol

(e)
y1, A  0.74
x1, A  0.4
F1  F2  VA  LA
F1  y1, AVA  x1, A LA
mol
VA  1.18
s
mol
LA  0.82
s

(f)
y1, B  0.9
x1, B  0.62
VA  VB  LB
y1, AVA  y1, BVB  x1, B LB
mol
VB  0.51
s
mol
LB  0.67
s

57
8.43

exp

B RT RT
1 ⟹ B 2 B B

RT ln
1 , , ,

RT RT
B B B
V V
RT RT
B 2 B
1 , ,
V V

1 1
RT ln RT 2 B 2 B
V V

VP 1 P
ln ln 2 B 2 B
RT V RT

2 B 2 B
ln ln
RT

ln ln P P P ln (I)

1 (II)

guess → solve II for → solve RHS for 1 repeat

RHS
y1 v
(I)
0 1.96 10 ---
0.01 1.95 10 10.76
0.05 1.96 10 10.14
0.00001 1.96 10 3.99
0.001 8.59
0.0001 6.30
0.00005 5.61
0.00006 5.79
0.000052 5.64
58
From Table B.3

P 260 Pa 10 ∙ 0.018 1.8 10

P 100 10 Pa

LHS I: ln P P P 5.64

y1 = 0.000052

59
8.44

(b) 0.8

(d)

At azeotrope: 1.06 2.65

RT ln RT ln

J J
A 3630 mol A 3798 mol

J
3720 mol

(e) F 2 mol s

0.47

0.2

1.46 mol s
.
L
.

0.54 mol
.
V F
. .

, 0.63

60
, 0.23

0.47 0.23
V V 0.32 mol s
0.63 0.23

61
8.45

(a)
In order to be consistent with Henry’s law, the activity coefficient of species 1 should approach
one as the mole fraction of species 1 goes to zero. For the given equation

lim 1Henry 's  30.5 112   0


x10

1Henry' s 1

Indeed, the activity coefficient expression is consistent with Henry’s law. It is also clear from
the expression that the activity coefficient will always be greater than unity. Therefore, the
fugacity of the mixture is always greater than the ideal fugacity based on Henry’s law reference
state. This suggests that the tendency for the molecules to escape into the vapor phase (vaporize)
in the real system is greater than the tendency when only unlike interactions exist. The unlike
interactions are stronger.

(b)
First, start with the Gibbs-Duhem as follows

 ln  1Henry 's  ln  2LR


x1  x2 0
x2 x2

For the given activity coefficient expression

 ln  1Henry ' s
 61.0 x2
x2

Hence,

 ln  2LR
x1  61.0 x 2  x 2 0
x 2
 ln  2LR
 61x1  0
x 2

Separate variables and integrate:

ln  2LR  61x 2  30.5 x22  C

For the Lewis/Randall reference state  2LR  1 when x 2  1. Therefore,

62
0  61  30.5  C
C  30.5
and
ln  2LR  61x 2  30.5 x 22  30.5
ln  2LR  30.5 x12

(c)
First, solve the problem for the solute (species 1). For vapor-liquid equilibrium

fˆ1v  fˆ1l
y1 P  x1 1Henry ' s[ 1
(Note: The fugacity coefficient does not appear in the equation because we are assuming
the vapor behaves ideally.)

Therefore,

y1  1Henry ' s[ 1 [ 1
  exp 30.5 1  x22  
x1 P P

Now, consider species 2.

fˆ2v  fˆ2l
y 2 P  x 2 2LR P2sat

Therefore,

y2  2LR P2sat P2sat P2sat


   2
exp 30.5x1   exp 30.5 1 x2  
2

x2 P P P

(d)
The following must be true for an azeotrope to form

y1 y 2
 1
x1 x 2

Substitute the expressions from Part (c) into the above relationship:

[1 P sat
exp 30.5 1 x22   2 exp 30.5x12 
P P

63
[ 1 exp 30.5 1 x22   P2sat exp 30.5 1 x2  
2

Insert numerical values and solve for x2:

x2  0.86
 x1  0.14

Now, we can calculate the pressure.

P   2LR P2sat   0.02  bar  exp 30.5  0.14   0.011  bar 


2

P  1.1 kPa

64
8.46
Since the pressure is 1 bar, we can assume the gas is ideal. We will also assume the liquid is
ideal. We can find the solubility of oxygen (species 1) in methanol (species 2) by solving for x1
in the following equation.

P  x1[ 1  x 2 P2sat  x1[ 1  1  x1 P2sat

Using Antoine’s Equation, we find

P2sat  0.168 bar 

From Table 8.2

[ 1  3179.4 bar 

Now, we can solve for the mole fraction.

P  P2sat 1  0.168 bar


x1    0.000262
[ 1  P2sat 3179.4  0.168 bar

At 100 bar, the Henry’s constant will change. We calculate the new value as follows.

 d ln[ 1    RT dP
V1

 [1  4.510 5
ln  
 3179.4 bar  8.314  298.15
100 10 5 -110 5  Pa 

[ 1 100 bar   571.4 bar 

Now calculate the mole fraction as before

P  P2sat 1  0.168 bar


x1    0.00146
[ 1  P2sat 571.4  0.168 bar

65
8.47
Since the pressure is 1 bar, we can assume the vapor behaves ideally. Also, since the liquid mole
fraction of carbon dioxide is so low, we can assume the solution is ideal (Henry’s limit).
Therefore,

yCO2 P  xCO2 [ CO2

To calculate the left hand side of the equation, we do the following:

P  y H 2O P  yCO2 P
y CO2 P  P  y H 2O P

By equating fugacities, we also have

y H 2 O P  x H 2 O PHsatO
2

Therefore,

yCO2 P  P  x H 2 O PHsatO  101325 Pa  0.999745 31188 Pa   0.7014 bar 


2

Now we can calculate Henry’s constant:

yCO2 P
 2751 bar 
0.7014 bar
[ CO2  
xCO2 0.000255

66
8.48
The following is true:

  ln [ i  Vi

  P   RT
 T

Plot the data given in the problem statement:

ln Hi

P
From the slope we can find the desired quantity:

slope 0.007573
R 8.314
T 292.55

Vi  18.42 J/mol bar


3
0.0001842 m /mol

67
8.49
Use the following relationship

  ln [ O 2  H O  hOv
   2 2

  1 / T   R
 P

Therefore, the following plot should be created with a linear fit to the data.

Plot of Henry's Constant Data

4.5

4
y = -3500.6x + 15.194
3.5
ln [

2.5

2
0.0031 0.0032 0.0033 0.0034 0.0035

1/T (K‐1)

From the trendline, we have

H O  hOv
2 2
 3500.6 K 
R
  J   J 
H O2  hOv 2   3500.6  K  8.314    29104
  mol K   
mol 

68
8.50
First, format the top of the spreadsheet as follows
-1 -1
T (ºC) 10 R (J mol K ) 8.314 A (J/mol) Initial Guess
T (K) 283.15

Caculated using data given in the


Measured Data Calculated using "A"
problem statement
E E 2
x1 x2 y1 y2 P (Pa) 1 (J/mol) 2 (J/mol) g ,exp (J/molg ,calc (J/mol) (gE -gE,calc)

Measured Data section:


Simply copy the given data into the cells below the appropriate column. Note that

x2  1  x1
y2  1  y1

Calculations with Data Given in the Problem Statement section:


The next section contains values of the activity coefficients and excess Gibbs energy calculated
without using A. The activity coefficients are calculated using Equations 8.11 and 8.12:

y1P
1  where P1sat is equal to the pressure when x1  1
x1P1sat

Calculate the excess Gibbs energy:

E
g exp  RT x1 ln  1  x2 ln  2 

Calculated using “A” section:


We are fitting the two-suffix Margules equation; therefore,

E
gcalc  Ax1x2

The value of A is equal to an initial guess that you enter as shown in the spreadsheet above.
Calculating

g 
2
E
exp  gcalc
E

should be straightforward. After entering all of the available data and equations in the
spreadsheet, the value of A is found using a solver function which minimizes the sum of
g  by varying A.
2
E
exp  gcalc
E

69
-1 -1
T (ºC) 10 R (J mol K ) 8.314 A (J/mol) 1398.328
T (K) 283.15

Caculated using data given in the


Measured Data Calculated using "A"
problem statement
E E 2
x1 x2 y1 y2 P (Pa) 1 (J/mol) 2 (J/mol) g (J/mol) g ,calc (J/mol) (gE -gE,calc)
0 1 0 1 6344 1.00 0.00 0.00
0.061 0.939 0.0953 0.9047 6590 1.70 1.00 77.63921 80.09 6.03
0.2149 0.7851 0.271 0.729 6980 1.45 1.02 227.3157 235.92 74.09
0.3187 0.6813 0.36 0.64 7140 1.33 1.06 302.1422 303.62 2.18
0.432 0.568 0.4453 0.5547 7171 1.22 1.10 332.0154 343.12 123.22
0.5246 0.4754 0.5106 0.4894 7216 1.16 1.17 356.1797 348.74 55.41
0.6117 0.3883 0.5735 0.4265 7140 1.10 1.24 334.0458 332.14 3.65
0.7265 0.2735 0.6626 0.3374 6974 1.05 1.36 275.2791 277.84 6.58
0.804 0.196 0.7312 0.2688 6845 1.03 1.48 227.6582 220.35 53.35
0.883 0.117 0.82 0.18 6617 1.01 1.60 154.7184 144.46 105.18
0.8999 0.1001 0.8382 0.1618 6557 1.01 1.67 132.9141 125.96 48.34
1 0 1 0 6073 1.00 0.00 0.00
SUM 478.04

A  1398 J/mol

 
2
E
OFg E is equal to the sum of the entries in the gexp  gcalc
E
column.

OFg E  478.04 J 2 / mol2 

70
8.51
First, format the top of the spreadsheet as follows

Caculated using data given


Measured Data Calculated using "A"
in the problem statement

1 -1,calc)/1] + [(2


2

1 (J/mol) 2 (J/mol) 1,calc (J/mol) 2,calc (J/mol) 1 -1,calc)/1]


2
2 -2,calc)/2]
2
2,calc)/2]
2
x1 x2 y1 y2 P (Pa)

Measured Data section:


Simply copy the given data into the cells below the appropriate column. Note that

x2  1  x1
y2  1  y1

The next section contains values of the activity coefficients and excess Gibbs energy calculated
without using A. The activity coefficients are calculated using Equations 8.15 and 8.16:

y1P
1  where P1sat is equal to the pressure when x1  1
x1P1sat

Calculated Using “A” section:


To calculate ã 1calc and ã 2calc we use the following expressions

 Ax22   Ax12 
1calc  exp    2calc  exp  
 RT   RT 

The calculations for the remaining columns should be obvious. The value of A is found using a
solver function which minimizes the sum of the of the entries in the following column

 1  1calc    2   2calc 
2 2

   
 1    2 

by changing A. We obtain

 J 
A  1424 
 mol 

    calc      calc 
2 2

OF is equal to the sum of the entries in the  1 1    2 2  column.


 1    2 
OF  1.49  10 3

71
8.52
This problem can be solved by creating a spreadsheet. First, format the top of the spreadsheet as
follows

-1 -1
T (ºC) 60 R (J mol K ) 8.314 A (J/mol) Initial Guess
T (K) 333.15

Caculated using data given in the


Measured Data Calculated using "A"
problem statement
E E 2
x1 x2 y1 y2 P (Pa) 1 (J/mol) 2 (J/mol) g ,exp (J/molg ,calc (J/mol) (gE -gE,calc)

Measured Data section:


Simply copy the given data into the cells below the appropriate column. Note that

x2  1  x1
y2  1  y1

The next section contains values of the activity coefficients and excess Gibbs energy calculated
without using A. The activity coefficients are calculated using Equations 8.11 and 8.12:

y1 P
1  where P1sat is equal to the pressure when x1  1
x1 P1sat

Calculate the excess Gibbs energy:

E
gexp  RT x1 ln  1  x2 ln  2 

Calculated using “A” section:


We are fitting the two-suffix Margules equation; therefore,

E
gcalc  Ax1x2

The A value is equal to an initial guess that you enter as shown in the spreadsheet above.
Calculating

g 
2
E
exp  gcalc
E

should be straightforward. After entering all of the available data and equations in the
spreadsheet, the value of A is found using a solver function which minimizes the sum of the
g  column by varying A.
2
E
exp  gcalc
E

72
T (ºC) 60 R (J mol-1 8.314 A (J/mol) 1137.182971
T (K) 333.15

Caculated using data given in the


Measured Data problem statement Calculated using "A"
E E E E 2
x1 x2 y1 y2 P (Pa) 1 2 g (J/mol) g ,calc (J/mol) (g -g ,calc)
0 1 0 1 51857 1.00 0.00 0.00
0.0672 0.9328 0.0912 0.9088 53431 1.39 1.00 71.12 71.28 0.03
0.2261 0.7739 0.267 0.733 55939 1.27 1.02 193.60 198.98 28.93
0.3201 0.6799 0.3526 0.6474 56741 1.20 1.04 237.12 247.49 107.53
0.432 0.568 0.448 0.552 57527 1.14 1.08 278.31 279.04 0.53
0.5203 0.4797 0.5203 0.4797 57633 1.10 1.11 283.28 283.83 0.30
0.6029 0.3971 0.5895 0.4105 57432 1.08 1.14 271.11 272.25 1.31
0.7095 0.2905 0.677 0.323 56989 1.04 1.22 241.99 234.38 57.80
0.7952 0.2048 0.7563 0.2437 56095 1.02 1.29 191.67 185.20 41.84
0.8752 0.1248 0.8386 0.1614 54934 1.01 1.37 129.48 124.21 27.83
0.8932 0.1068 0.86 0.14 54629 1.01 1.38 114.76 108.48 39.49
1 0 1 0 52190 1.00 0.00 0.00
SUM 305.58

Therefore,

 J 
A  1137.2 
 mol 

In the example in the text, A was calculated to be 1143 J/mol, which is 0.51% larger than the
value obtained using the objective function. The value obtained using the objective function is
more accurate because it utilizes more empirical data.

73
8.53

(a)
We will find the 3-suffix Margules parameters using a spreadsheet. First, format the top of the
spreadsheet as follows:

T 40 ºC R 8.314 J mol-1 K-1 A Guess J/mol


T 313.15 K B Guess J/mol

Calculated using 3-Suffix


Measured Data
Margules Equation
x1 x2 y1 y2 P (Pa) 1 2 Pcalc (Pa) (P-Pcalc)2

Measured Data section:


Copy the given data into the cells below the appropriate column. Note that

x2  1  x1
y2  1  y1

Calculated using 3-Suffix Margules Equation section:


Reference the cells containing the A and B parameters to calculate the activity coefficients using

RT ln  1   A  3B  x22  4Bx23
RT ln  2   A  3B  x12  4Bx13

The pressure is then calculated as follows:

P  x1 1P1sat  x2 2 P2sat

where

P1sat  P when x1  1
P2sat  P when x1  0

The last column is self-explanatory. Use the solver function to simultaneously find the values of
A and B that minimize the sum of (P-Pcalc)2. Doing so, we obtain

A  1335 J/mol 
B  244 J/mol 

(b)
Format the spreadsheet as follows

74
T 40 ºC R 8.314 J mol-1 K-1 A Guess J/mol
T 313.15 K B Guess J/mol

Caculated using data given Calculated using "A"


Measured Data
in the problem statement and "B"
gE gEcalc
x1 x2 y1 y2 P (Pa) 1 2 (gE- gE calc)2
(J/mol) (J/mol)

Measured Data section:


Copy the given data into the cells below the appropriate column. Note that

x2  1  x1
y2  1  y1

Calculations with Data Given in the Problem Statement section:


This section contains values of the activity coefficients and excess Gibbs energy calculated
without using A and B. The activity coefficients are calculated using Equations 8.11 and 8.12:

y1 P
1  where P1sat is equal to the pressure when x1  1
x1 P1sat

Calculate the excess Gibbs energy using

E
g exp  RT x1 ln  1  x2 ln  2 

Calculated using “A” and “B” section:


We are fitting the two-suffix Margules equation; therefore,

E
g calc  x1 x 2 A  Bx1  x 2 

Calculating

g 
2
E
 gcalc
E

should be straightforward. The values of A and B are found using a solver function which
 
2
minimizes the sum of the entries in the g E  gcalc
E
column. We obtain

A  1364 J/mol
B  247.5 J/mol

(c)
Format the spreadsheet as follows:

75
-1 -1
T 40 ºC R 8.314 J mol K A Guess J/mol
T 313.15 K B Guess J/mol

Caculated using data given


Measured Data Calculated using "A" and "B"
in the problem statement
1 -1,calc)/1] +
2
1 2 1,calc  2,calc  1 -1,calc)/1]  2 -2,calc)/2] 
2 2
x1 x2 y1 y2 P (Pa)
[(2 2,calc)/2] 
2

Measured Data section:


Simply copy the given data into the cells below the appropriate column. Note that

x2  1  x1
y2  1  y1

The next section contains values of the activity coefficients and excess Gibbs energy calculated
without using A and B. The activity coefficients are calculated using Equations 8.15 and 8.16:

y1P
1  where P1sat is equal to the pressure when x1  1
x1P1sat

Calculated Using “A” and “B” section:


To calculate  1 and  2 we use the following expressions

RT ln 1calc   A  3B x22  4Bx23


RT ln  2calc   A  3B  x12  4Bx13

The calculations for the remaining columns should be obvious. The values of A and B are found
using a solver function which minimizes the sum of the entries in the following column:

 1  1calc    2   2calc 
2 2

   
 1    2 

Doing so, we obtain

A  1369 J/mol
B  258 J/mol

76
(d)
Using the excess Gibbs energy calculated in Part (b), create the following spreadsheet:

x1-x2 gE/x1*x2
-1
-0.9 1129.969
-0.8 1206.466
-0.7 1205.426
-0.6 1243.795
-0.5 1247.326
-0.4 1259.465
-0.3 1274.243
-0.2 1296.296
-0.1 1328.461
0 1353.076
0.1 1447.655
0.2 1398.372
0.3 1414.083
0.4 1459.965
0.5 1494.193
0.6 1527.094
0.7 1526.423
0.8 1601.432
0.9 1647.631
1

gE
If we plot of versus x1  x 2 , the slope of the resulting line is B and the intercept is A.
x1 x 2

E
g /x1x2 vs. x1-x2
1700
1600
gE/x1x 2 (J/mol)

1500
1400
1300
y = 255.17x + 1371.7
1200
R2 = 0.9683
1100
1000
-1 -0.5 0 0.5 1
x 1-x 2

77
As you can see from the graph,

A  1372 J/mol 
B  255 J/mol 

The answers obtained in parts A – D agree well with each other. The average value of A is 1360
J/mol, and the standard deviation of the answers is 17 J/mol. The average value of B is 251
J/mol, and the standard deviation is 6.5 J/mol.

From ThermoSolver, we obtain the following values:

Objective Function A (J/mol) B (J/mol)


Pressure 1335 244
Gibb’s Energy 1364.5 247.5
Activity Coefficient 1368.6 258.6

These values are equal to those obtained in Parts (a) – (c).

78
8.54
To test for thermodynamic consistency, we use:

1
 1 
 ln  
0
 dx1  0
2 

However, the value of the above integral is not useful unless it is referenced to its absolute value.
We can calculate the following
1
 1 
0
 ln  
2
 dx1

1
 
0 ln   12  dx1

and confirm that it is approximately equal to zero. We can obtain the values of the activity
coefficients from experimental data using the following equations, which assume ideal behavior
of the gas phase:

y1 P
y1P  x1 1P1sat  1 
x1 P1sat

y 2 P  x2 2 P2sat  2 
1  y1  P
1  x1  P2sat
To find the saturation pressures from the provided data, we must realize

P1sat  Px1  1
P2sat  Px1  0

With the above expressions, the following table was made in created in a spreadsheet

x1 1 2 ln(1/2)


0 1
0.05 1.501 1.000 0.406
0.1 1.455 1.005 0.371
0.15 1.413 1.008 0.338
0.2 1.370 1.017 0.298
0.25 1.329 1.025 0.260
0.3 1.293 1.035 0.223
0.35 1.257 1.049 0.181
0.4 1.223 1.067 0.137

79
0.45 1.191 1.090 0.088
0.5 1.161 1.117 0.038
0.55 1.134 1.147 -0.011
0.6 1.109 1.183 -0.065
0.65 1.086 1.221 -0.118
0.7 1.064 1.281 -0.185
0.75 1.046 1.345 -0.252
0.8 1.03 1.420 -0.321
0.85 1.018 1.489 -0.380
0.9 1.008 1.626 -0.479
0.95 1.002 1.77 -0.569
1 1

Use the spreadsheet to create the following graph:

ln( 1/ 2) vs. x 1


0.6

0.4

0.2
ln(  1 / 2 )

0
0 0.2 0.4 0.6 0.8 1
-0.2

-0.4
2
ln( 1/ 2) = -0.5924x - 0.4573x + 0.419
-0.6 2
R = 0.9992
-0.8
x1

Integrate the trendline provided in the plot:

 -0.5924 x12 - 0.4573x1 + 0.419  dx1


0
1
 0.027

2
-0.5924 x - 0.4573x1 + 0.419 dx1
1
0

This integral is approximately zero, so the data are thermodynamically consistent.

80
8.55
To test for thermodynamic consistency, we use:

1
 
 ln  12 dx1  0
0

However, the value of the above integral should be referenced to its absolute value.
We can obtain the values of the activity coefficients from experimental data using the following
equations, which assume ideal behavior of the vapor phase:

y1P
y1P  x1 1P1sat  1 
x1P1sat

y 2 P  x2 2 P2sat  2 
1  y1 P
1  x1 P2sat
To find the saturation pressures from the provided data, we must realize

P1sat  Px1  1
P2sat  Px1  0

With the above expressions, the following table was made in created in a spreadsheet

x1 1 2 ln(1/2)


0 1
0.0821 0.494107 0.986964 -0.69188
0.1953 0.569841 0.950838 -0.51199
0.2003 0.542755 0.957594 -0.56777
0.3365 0.681444 0.875216 -0.25026
0.4182 0.753447 0.820105 -0.08477
0.4917 0.8122 0.774084 0.048066
0.595 0.888038 0.715016 0.21671
0.709 0.942786 0.649849 0.3721
0.8182 0.980142 0.595358 0.498534
0.8768 0.949396 0.91193 0.040263
0.938 0.92917 1.09213 -0.16159
0.972 0.963997 1.133066 -0.16159
1 1

Use the spreadsheet to create the following graph:

81
ln( 1/ 2) vs. x 1
0.6

0.4

0.2

ln(  1 /  2)
0

-0.2

-0.4

-0.6

-0.8
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x1

Clearly, the area above the line y=0 is less than the area below. The value of the following
integral can be estimated graphically.

1
 1 
 ln  2 dx1
0
1
 1 
 ln  2  dx1
0

There are a number of ways to do this. One possibility is to print the graph. The area above the
line y=0 is cut out, as is the area below. The two pieces of paper are weighed. Assuming a
constant density of the printer paper, the weights should be equal. If not, the areas are different;
thus, the data are not thermodynamically consistent. The following data are obtained for the
above graph

Mass (milligrams)
Area above y=0 137.5
Area below y=0 260.4
Total area 397.9

It should be noted that the trends in ln 1 /  2  were extended to x1  0 and x1  1 to estimate the
entire areas above and below the x-axis. Using these data, we estimate

1
ã 
 ln ã 12 dx1 137.5  260.4
0   0.309
1
ã  397.9
 ln ã 12  dx1
0

Therefore, the data are not thermodynamically consistent.

82
8.56
The data are thermodynamically consistent if:

1
 
 ln  12 dx1  0
0

To find the activity coefficients, we can use the following equations:

1 
y1P
2 
1  y1 P
x1P1sat 1  x1 P2sat
The pressure is constant at 1 atm (1.013 bar), but the saturation pressures are dependent upon
temperature. We can find water’s saturation pressures from the steam tables. Use Antoine’s
equation data in Appendix A to calculate the sat. pressure of acetone. We can create the
following table.

x1 y1 T (ºC) P1sat (bar) P2sat (bar) 1  ln(1/2)


0 0 100 3.711 1.013 1.00
0.015 0.325 89.6 2.812 0.701 7.81 0.99 2.07
0.036 0.564 79.4 2.104 0.463 7.54 0.99 2.03
0.074 0.734 68.3 1.502 0.290 6.69 1.00 1.90
0.175 0.8 63.7 1.296 0.237 3.57 1.04 1.24
0.259 0.831 61.1 1.190 0.211 2.73 1.09 0.91
0.377 0.84 60.5 1.167 0.205 1.93 1.27 0.42
0.505 0.849 59.9 1.144 0.199 1.49 1.55 -0.0422
0.61 0.868 59 1.110 0.191 1.30 1.80 -0.32
0.804 0.902 58.1 1.077 0.184 1.06 2.75 -0.96
0.899 0.938 57.4 1.052 0.178 1.00 3.49 -1.25
1 1 56.3 1.013 0.168 1.00
Note: The last row of data (x1=1) was obtained by finding the temperature that results in a
saturation pressure of 1 atm in Antoine’s equation.

83
Now, create the following plot.

ln( 1/ 2) vs. x 1

2.5
2
1.5
1
ln(  1 / 2 )

0.5
0
-0.5
-1
-1.5
0 0.2 0.4 0.6 0.8 1
x1

Clearly, the area above the x-axis is greater than the area below. The value of the following
integral can be estimated graphically.

1
 1 
 ln  2 dx1
0
1
 1 
 ln  2  dx1
0

There are a number of ways to do this. One possibility is to print the graph. The area above line
y=0 is cut out, as is the area below. The two pieces of paper are weighed. Assuming a constant
density and thickness of the printer paper, the weights should be equal. If not, the areas are
different; thus, the data are not thermodynamically consistent. The following data are obtained
for the above graph

Mass (milligrams)
Area above y=0 210.0
Area below y=0 159.7
Total area 369.7

 
It should be noted that the trends in ln  1  were extended to x1  0 and x1  1 to estimate the
2 
entire areas above and below the x-axis. Using these data, we estimate

84
1
 ã1 
  ã 2 dx1
ln
210.0  159.7
0   0.1361
1
ã  369.7
 ln ã 12  dx1
0

Therefore, the data are not thermodynamically consistent.

85
8.57
Let “1” designate benzene and “2” designate isooctane. To solve this problem, we will need to
know molar volumes and saturation pressures. First, calculate these quantities from given data.

MWi  m3   m3 
vi  v1  8.94  10  5   v 2  1.66  10  4  
i  mol   mol 

 B 
P1sat  14.2 bar P2sat  8.86 bar
Pisat  exp  A 
 C  T 

In order to find the constants A and B, we need values for both activity coefficients. The solution
method for finding the activity coefficient of benzene will be shown, and then the activity
coefficient of isooctane can be found analogously. First, equate fugacities:

y1ˆ1 P  x1 1 f1o

The reference fugacity is defined as follows

 P vl 
f1o  f1l  P1sat 1sat exp   1 dP 
 sat RT 
 P1 

Find the saturation fugacity coefficient from the following relationship:

P1sat
 z1 1 dP P 1 B' P  1 dP  P
sat sat

ln      B'dP
1 1
sat
1
Plow P Plow P Plow

where

B
B'   1.2510 7 Pa -1 
RT

If we integrate the expression and let Plow go to zero, we obtain

1sat  0.837

Therefore,

11.6105 Pa 8.94 10 5 


f  14.2 bar   0.837 exp  
o
dP   11.82  bar 
14.2105 Pa 8.314  473.15 
1

We also need the fugacity coefficient of species 1 in the mixture.

86
  
P
Z1 1
ln ˆ1  dP
Plow P

where

 
Z1  n 1 Bmix
'
P 
n1 
and
'
Bmix  y12 B11
' '
 2 y1 y 2 B12  y 22 B22

Therefore,

Z1  1  2y1  y12  B11'   2y2  2y1 y2  B12'  y22 B22


' 
P
and

   2y  y  B
P
' 
ln ˆ1  1
2
1
'
11   2y2  2y1 y2  B12'  y22 B22  dP
Plow

ˆ1  0.870
(Note: When you calculate B12' using a geometric average, its value is negative, even

though the product of two negatives is positive.)

Now we can calculate the activity coefficient.

1 
y1ˆ1 P

0.37 0.87 11.6 bar   1.26
x1 f1o 0.2511.82 bar 

From a similar method

 2  1.03

We have two equations with two unknowns: A and B.

RT ln 1   A  3B x22  4Bx23 RT ln  2   A  3B x12  4Bx13

Solve simultaneously:

 J   J 
A  1616   B  122  
 mol   mol 

87
(b)
A mixture of benzene and isooctane will split into two partially miscible phases if

2g
0
x12

The expression for Gibbs energy is

g  x1 g1  x2 g 2  RT x1 ln x1  x2 ln x2  x1 x 2 A  Bx1  x2 

Therefore,

2 g 1 1
 RT     2A 12Bx1  6B
x12  x1 x2 

Assume that A and B are independent of temperature. For the mixture to split into two phases

2A 12Bx1  6B
T
1 1
R  
 x1 x2 

T  81.1 K

The freezing points of benzene and isooctane are 278 K and 165 K, respectively. It is unlikely
this mixture forms a eutectic at this low a temperature, so it will solidify before it reaches the
limit of liquid instability.

88
8.58
For a pure species:
iv  il

For the Peng-Robinson equation of state:


RT ai i
P  2
vi  bi vi  2bi vi  bi2
where
0.45724 R 2Tc2,i
  
2
ai  , i  1  0.37464  1.54226i  0.26992i2 1  Tr0.5 
,i  , and
P c ,i

0.07780RTc,i
bi 
Pc,i
The values of the critical properties from Appendix A.1and the resulting calculated parameters,
ai, i, and bi are:

ai bi
Tc Pc  Jm 3   m3 
Species  
[K] [bar]  mol 2   mol 
   
n-C5H12 469.6 33.74 0.251 2.066 1.321 9.00 x 10-5

Knowing the Peng-Robinson parameters, we can calculate the molar volume, vi. When the
solution has three real roots, we assign the smallest root to the liquid phase and the largest root to
the vapor phase. These values can then be used to find, the fugacity coefficient of the liquid and
vapor, respectively, using the expression:

 v  b  P 
ln i  zi  1  ln  i i 
 a i 

 vi  1  2 bi 


 ln
 RT  2 2bi RT  v
 i
 1  2 bi
 
To find the saturation pressure at a given temperature, we solve for liquid and vapor volumes and
then use those values for the fugacity coefficient. This process requires iterative solution and the
saturation pressure is found as the pressure where the fugacity of the vapor and liquid are equal.
The values that satisfy this criteria

vil , viv i Pi sat


Pi sat
Species  m3   m3  when
Antoine error
 mol   mol    il ,
v [bar]
[bar]
    i

-4
n-C5H12 1.46 x 10 2.47 x 10-3 0. 841 10.45 10.31 1.4%

89
8.59
For a pure species:
iv  il

For the Peng-Robinson equation of state:


RT ai i
P  2
vi  bi vi  2bi vi  bi2
where
0.45724 R 2Tc2,i
  
2
ai  , i  1  0.37464  1.54226i  0.26992i2 1  Tr0.5 
,i  , and
P c ,i

0.07780RTc,i
bi 
Pc,i
The values of the critical properties from Appendix A.1and the resulting calculated parameters,
ai, i, and bi are:

ai bi
Tc Pc  Jm 3   m3 
Species  
[K] [bar]  mol 2   mol 
   
C3H8 370 42.44 0.152 1.02 1.177 5.64 x 10-5

Knowing the Peng-Robinson parameters, we can calculate the molar volume, vi. When the
solution has three real roots, we assign the smallest root to the liquid phase and the largest root to
the vapor phase. These values can then be used to find, the fugacity coefficient of the liquid and
vapor, respectively, using the expression:

 v  b  P 
ln i  zi  1  ln  i i 
 a i 

 vi  1  2 bi 


 ln
 RT  2 2bi RT  v
 i
 1  2 bi
 
To find the saturation pressure at a given temperature, we solve for liquid and vapor volumes and
then use those values for the fugacity coefficient. This process requires iterative solution and the
saturation pressure is found as the pressure where the fugacity of the vapor and liquid are equal.
The values that satisfy this criteria

vil , viv i Pi sat


Pi sat
Species  m3   m3  when
Antoine error
 mol   mol    il ,
v [bar]
[bar]
    i

-5
C3H8 7.89 x 10 4.30 x 10-3 0.903 4.71 4.74 0.6%

90
8.60
For a pure species:
iv  il

For the Peng-Robinson equation of state:


RT ai i
P  2
vi  bi vi  2bi vi  bi2
where
0.45724 R 2Tc2,i
  
2
ai  , i  1  0.37464  1.54226i  0.26992i2 1  Tr0.5 
,i  , and
P c ,i

0.07780RTc,i
bi 
Pc,i
The values of the critical properties from Appendix A.1and the resulting calculated parameters,
ai, i, and bi are:

ai bi
Tc Pc  Jm 3   m3 
Species  
[K] [bar]  mol 2   mol 
   
C6H6 562.1 48.94 0.212 2.041 1.4613 7.43 x 10-5

Knowing the Peng-Robinson parameters, we can calculate the molar volume, vi. When the
solution has three real roots, we assign the smallest root to the liquid phase and the largest root to
the vapor phase. These values can then be used to find, the fugacity coefficient of the liquid and
vapor, respectively, using the expression:

 v  b  P 
ln i  zi  1  ln  i i 
 a i 

 vi  1  2 bi 


 ln
 RT  2 2bi RT  v
 i
 1  2 bi
 
To find the saturation pressure at a given temperature, we solve for liquid and vapor volumes and
then use those values for the fugacity coefficient. This process requires iterative solution and the
saturation pressure is found as the pressure where the fugacity of the vapor and liquid are equal.
The values that satisfy this criteria

vil , viv i Pi sat


Pi sat
Species  m3   m3  when
Antoine error
 mol   mol    il ,
v [bar]
[bar]
    i

-5
C6H6 9.98x 10 8.81 x 10-3 0.929 3.49 3.52 0.8%

91
8.61
Following Example 8.15, we get:

T, xi

Ai  Bi 
Antoine Pi sat  exp  A1 
Bi 
Coefficients  T  Ci 
Ci
P   xi Pi sat Initial Estimate
Raoult’s Law
xi Pi sat
yi 
P

P (1) , yi(1)

Pi sat ai bi
Species
xi Ai Bi Ci Antoine Tc,i Pc,i  Jm  3
 m3 
[bar] [K] [bar]  mol2   mol 
   
methane (1) 0.2 8.6041 897.84 -7.16 283.8 190.6 46.0 0.2303 4.31 x 10-5
n-pentane (2) 0.8 9.2131 2477.1 -39.94 1.075 469.6 33.74 1.906 1.45 x 10-4

EOS Parameters
Tc,i ai a  y12 a1  2 y1 y2 a1a2  y22 a2
Pc,i bi b  y1b1  y2b2
RT a
P  2  v v , vl
v b v
m
2 yk aik
bi  v  b P 
ln ˆi   ln  
k 1
 ˆiv , ˆil
v b  RT  RTv
xiˆil
yi 
ˆiv

No P
 j 1
 P  j   yi j 
y i  1?

Yes

P, yi

92
av bv vv vl
n  Jm3 
 
 mol2 
 m3 
 
 mol 
 m3 
 
 mol 
 m3 
 
 mol 
ˆ1v ˆ2v ˆ1l ˆ2l y1 y2 y i
P(k+1)
[bar]

1 0.252 4.55 10-5 0.000346 0.000176 0.90 0.49 1.63 0.14 0.36 0.22 0.59 71.16

2 0.313 5.49 10-5 0.000631 0.000183 0.94 0.68 2.43 0.19 0.52 0.23 0.74 41.63

3 0.470 6.72 10-5 0.000838 0.000187 0.96 0.71 3.09 0.24 0.64 0.27 0.92 30.90

4 0.568 7.41 10-5 0.000868 0.000188 0.98 0.67 3.32 0.26 0.68 0.31 0.99 28.34

5 0.600 7.63 10-5 0.000843 0.000188 1.00 0.64 3.35 0.26 0.67 0.33 1.00 28.03

93
8.62
Following Example 8.16, we get:
T, xi

Ai  Bi 
Antoine Pi sat  exp  A1 
Bi 
Coefficients  T  Ci 
Ci
P   xi Pi sat Initial Estimate
Raoult’s Law
xi Pi sat
yi 
P

P (1) , yi(1)

Pi sat ai bi
Species
xi Ai Bi Ci Antoine Tc,i Pc,i  Jm  3
 m3 
[bar] [K] [bar]  mol2   mol 
   
methane (1) 0.2 8.6041 897.84 -7.16 283.8 190.6 46.0 0.191 2.68 x 10-5
n-pentane (2) 0.8 9.2131 2477.1 -39.94 1.075 469.6 33.74 2.58 9.00 x 10-5

EOS Parameters

a  y12  a 1  2 y1 y2  a 1  a 2 1  k12   y22  a 2


Tc,i ai
Pc,I i
 bi b  y1b1  y2b2
RT a
P  2  v v , vl
v  b v  2bv  b 2

b  v  b P 
ln ˆi  i  z  1  ln  
b  RT 

a  bi 2 m
 
 v  1 2 b   
   k ik   v  1  2 b   ˆiv ,ˆil
y  a  ln

2 2bRT  b a k 1
   
x ˆ l
yi  i v i
ˆi

No P
 j 1
 P
j
 y 
i
j

y i  1?

Yes

P, yi

94
av bv vv vl
n  Jm3 
 
 mol2 
 m3 
 
 mol 
 m3 
 
 mol 
 m3 
 
 mol 
ˆ1v ˆ2v ˆ1l ˆ2l y1 y2 y i
P(k+1)
[bar]

1 0.216 2.83 10-5 0.000348 0.000108 0.91 0.41 3.11 0.04 0.69 0.07 0.76 71.16

2 0.227 2.88 10-5 0.000479 0.000109 0.93 0.54 3.89 0.05 0.84 0.07 0.91 54.10

3 0.268 3.12 10-5 0.000515 0.000109 0.93 0.52 4.22 0.05 0.90 0.08 0.98 49.23

4 0.281 3.20 10-5 0.000514 0.000109 0.94 0.50 4.29 0.05 0.91 0.08 1.00 48.33

5 0.285 3.21 10-5 0.000511 0.000109 0.94 0.49 4.30 0.05 0.91 0.08 1.00 48.19

6 0.285 3.22 10-5 0.000511 0.000109 0.94 0.48 4.30 0.05 0.91 0.09 1.00 48.16

7 0.286 3.22 10-5 0.00051 0.000109 0.94 0.48 4.30 0.05 0.91 0.09 1.00 48.15

95
8.63
Following Example 8.15, we get:
T, xi

Ai  Bi 
Antoine Pi sat  exp  A1 
Bi 
Coefficients  T  Ci 
Ci
P   xi Pi sat Initial Estimate
Raoult’s Law
xi Pi sat
yi 
P

P (1) , yi(1)

Pi sat ai bi
Species
xi Ai Bi Ci Antoine Tc,i Pc,i  Jm  3
 m3 
[bar] [K] [bar]  mol2   mol 
   
Carbon
0.3 15.97 3103.4 -0.16 2099 304.2 73.76 0.366 4.29 x 10-5
Dioxide (1)
Benzene (2) 0.7 9.2806 2788.5 -52.36 1.80 562.1 48.94 1.88 1.19 x 10-4

EOS Parameters
Tc,i ai a  y12 a1  2 y1 y2 a1a2  y22 a2
Pc,i bi b  y1b1  y2b2
RT a
P  2  v v , vl
v b v
m
2 yk aik
bi  v  b P 
ln ˆi   ln  
k 1
 ˆiv , ˆil
v b  RT  RTv
xiˆil
yi 
ˆiv

 j 1
 P    yi 
j j
No P
y i  1?

Yes

P, yi

96
av bv vv vl
n  Jm3 
 
 mol2 
 m3 
 
 mol 
 m3 
 
 mol 
 m3 
 
 mol 
ˆ1v ˆ2v ˆ1l ˆ2l y1 y2 y i
P(k+1)
[bar]

1 1464 2.71 10-3 0.00273 0.000147 12.78 1.62 6.93 0.67 0.16 0.29 0.45 10.00

2 17.47 3.19 10-4 0.006809 0.000148 0.99 0.97 14.94 1.45 4.53 1.05 5.58 4.53

3 0.597 6.52 10-5 0.00034 0.000144 4.21 0.43 2.91 0.29 0.21 0.47 0.68 25.28

4 1.605 9.76 10-5 0.001673 0.000145 0.96 0.80 4.15 0.41 1.29 0.35 1.65 17.16

5 1.180 9.19 10-5 0.000129 0.000143 3.01 0.27 2.63 0.26 0.26 0.68 0.94 28.24

6 0.823 7.15 10-5 0.00073 0.000143 1.10 0.59 2.79 0.28 0.76 0.33 1.09 26.48

7 0.761 6.81 10-5 0.000837 0.000143 0.98 0.63 2.59 0.26 0.79 0.29 1.08 28.77

8 0.689 6.49 10-5 0.00078 0.000142 0.96 0.62 2.41 0.24 0.75 0.27 1.03 31.08

9 0.662 6.35 10-5 0.000786 0.000142 0.95 0.63 2.36 0.24 0.75 0.26 1.01 31.87

10 0.651 6.29 10-5 0.000787 0.000142 0.94 0.63 2.34 0.23 0.74 0.26 1.00 32.19

11 0.647 6.27 10-5 0.000788 0.000142 0.94 0.63 2.33 0.23 0.74 0.26 1.00 32.30

12 0.645 6.26 10-5 0.000789 0.000142 0.94 0.63 2.33 0.23 0.74 0.26 1.00 32.34

13 0.645 6.26 10-5 0.000789 0.000142 0.94 0.63 2.33 0.23 0.74 0.26 1.00 32.35

14 0.645 6.26 10-5 0.000789 0.000142 0.94 0.63 2.33 0.23 0.74 0.26 1.00 32.35

97
8.64
Following Example 8.16, we get:
T, xi

Ai  Bi 
Antoine Pi sat  exp  A1 
Bi 
Coefficients  T  Ci 
Ci
P   xi Pi sat Initial Estimate
Raoult’s Law
xi Pi sat
yi 
P

P (1) , yi(1)

Pi sat ai bi
Species
xi Ai Bi Ci Antoine Tc,i Pc,i  Jm  3
 m3 
[bar] [K] [bar]  mol2   mol 
   
Carbon
0.3 15.97 3103.4 -0.16 2099 304.2 73.76 0.338 2.67 x 10-5
Dioxide (1)
Benzene (2) 0.7 9.2806 2788.5 -52.36 1.80 562.1 48.94 2.59 7.43 x 10-5

EOS Parameters

a  y12  a 1  2 y1 y2  a 1  a 2 1  k12   y22  a 2


Tc,i ai
Pc,I i
 bi b  y1b1  y2b2
RT a
P  2  v v , vl
v  b v  2bv  b 2
b  v  b P 
ln ˆi  i  z  1  ln  
b  RT 

a  bi 2 m
 
 v  1 2 b   
   k ik   v  1  2 b   ˆiv ,ˆil
y  a  ln

2 2bRT  b a k 1
   
x ˆ l
yi  i v i
ˆi

No P
 j 1
 P  j   yi j 
y i  1?

Yes

P, yi

98
av bv vv vl
n  Jm3 
 
 mol2 
 m3 
 
 mol 
 m3 
 
 mol 
 m3 
 
 mol 
ˆ1v ˆ2v ˆ1l ˆ2l y1 y2 y i
P(k+1)
[bar]

1 1356 1.69 10-3 0.001703 8.62 10-5 0.86 0.37 1.30 0.10 0.45 0.18 0.634 54.5

2 37.5 2.82 10-4 0.007119 8.65 10-5 0.91 0.56 1.78 0.12 0.59 0.15 0.739 40.3

3 0.33 2.67 10-5 0.000297 8.47 10-5 0.93 0.63 2.24 0.15 0.72 0.16 0.887 35.7

4 1.67 5.97 10-5 0.001284 8.56 10-5 0.94 0.61 2.47 0.16 0.79 0.19 0.971 34.7

5 0.62 3.69 10-5 8.38 10-5 8.46 10-5 0.96 0.57 2.53 0.16 0.79 0.20 0.995 34.5

6 0.81 4.18 10-5 0.000707 8.51 10-5 0.96 0.56 2.54 0.16 0.79 0.21 0.999 34.5

7 0.58 3.56 10-5 9.40 10-5 8.45 10-5 18.5 0.63 14.10 0.18 0.23 0.20 0.43 10.00

8 0.66 3.78 10-5 0.000606 8.49 10-5 0.99 0.96 32.13 0.42 9.73 0.31 10.04 4.33

9 0.47 3.17 10-5 0.000383 8.44 10-5 4.09 0.12 3.49 0.05 0.26 0.27 0.52 43.47

10 0.45 3.11 10-5 0.000449 8.43 10-5 0.95 0.74 6.37 0.08 2.02 0.08 2.10 22.78

11 0.42 3.01 10-5 0.000427 8.42 10-5 2.60 0.08 3.21 0.04 0.37 0.36 0.73 47.75

12 0.41 2.97 10-5 0.000429 8.42 10-5 0.95 0.51 4.26 0.06 1.35 0.08 1.43 35.01

13 0.41 2.96 10-5 0.000429 8.42 10-5 1.80 0.10 3.09 0.04 0.51 0.29 0.81 49.92

14 0.41 2.95 10-5 0.000429 8.42 10-5 0.93 0.47 3.74 0.05 1.21 0.07 1.28 40.33

15 0.41 2.95 10-5 0.000429 8.42 10-5 0.93 0.35 2.99 0.04 0.96 0.08 1.04 51.74

16 0.41 2.95 10-5 0.00043 8.42 10-5 0.88 0.42 2.88 0.04 0.99 0.06 1.05 54.00

99
8.65

100
x1
90 y1

80

70

60
P [bar]

50

40

30

20

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x1, y1

100
8.66
(a)

100
x1
90 y1

80

70

60
P [bar]

50

40

30

20

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x1, y1

(b)

100
x1
90 y1

80

70

60
P [bar]

50

40

30

20

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x1, y1

101
8.67

100
x1
90 y1

80

70

60
P [bar]

50

40

30

20

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x1, y1

102
8.68
(a)

100
x1
90 y1

80

70

60
P [bar]

50

40

30

20

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x1, y1

(b)

100
x1
90 y1

80

70

60
P [bar]

50

40

30

20

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x1, y1

103
8.69
Use the given activity coefficients at infinite dilution to obtain values for A and B in the three-
suffix Margules equation:

RT ln  a   A  3B xb2  4Bxb3  RT ln  a  A  B


RT ln  b   A  3B  xa2  4Bxa3  RT ln  b  A  B

Substitute values:

  J 
8.314    300 K   ln8  A  B
  mol  K 
  J 
8.314    300 K   ln15  A  B
  mol K 

 J   J 
A  5970  B  784 
 mol   mol 

After noting that

xa  xb  1
xa  xb  1

set the liquid fugacities equal:

xa  a  xa  a
  3   3

xa exp 
  A  3B  1 x a 
 2
 4B 1 x a    x exp 
  A  3B  1 x a 
 2
 4B 1 x a 
 RT  a
 RT 
   
 
x    x 
b b b b
  A  3B   x 2  4 B  x 3    A  3 B   x   2  4 B  x  3 
1  xa  exp  a

RT
a
  1  xa  exp 
 
a

RT
a


   

Determine the values of x 


a and xa using a solver function on mathematical software.

x a  0.89
x a  0.23

104
Therefore,

x b  0.11
x b  0.77

105
8.70
To determine which species is on top, we must calculate the density of both species.

MWi
i 
vi
86  g/mol
1   0.659 g/cm 3 
 3
130.5cm / mol
58  g/mol
2   0.79 g/cm 3 
73.4 cm / mol
3

Since phase  has more of the less dense component (hexane), it will be on top. Therefore, the
schematic should look similar to the following:

T = 15 oC P = 300 bar
liquid
n = 10 mol 

liquid

n = 20 mol

(b)
Because the mixture separates into two phases, it suggests that the species prefer like-like
interactions. Therefore, the like interactions are stronger.

(c)
We can calculate the pure species fugacity with the following equation:

 P v1sat 
f1l  1sat P1sat exp   dP 
 sat RT 
 P1 

Since the saturation pressure is low, we can assume the fugacity coefficient is one.
We can also assume that the molar volume is independent of pressure.

f1l  65.0 kPa 

(d)
For phase equilibrium of species 1 to exist, the following must be true:

106
x1  1 P1sat  x1  1 P1sat
 x1  1  x1  1

For the two suffix Margules equation, we can obtain expressions for the activity coefficients.

 A x 2   A x 2 
1  exp 
  2  1  exp 
  2 
 RT   RT 
   

Substitute the above expressions into the expression that equates fugacities:

  2   2

 x1 exp 
A  x 2    x  exp  A  x 2 
 RT  1
 RT 
   

Now, we can solve for A by substituting known values. We obtain:

 J 
A  5535 
 mol 

(e)
The lowest temperature at which the two species are completely miscible occurs when

RT
 2A
x1 x 2

If we assume that A is independent of temperature, then we can solve for the temperature easily.

2 Ax1 x2
T
R

We need to solve for the mole fractions of a completely miscible solution.

0.220 mol 0.810 mol


x1   0.4
30 mol
x2  0.6

Now substitute the known quantities:

107
  J 
2  5535    0.4  0.6
  mol 
T  320  K 
  J 
8.314  
  mol K 

(f)
We can calculate Henry’s constant from the following equation:

[1
 1 
f1

Calculate the activity coefficient:

 
 5535  J/mol 1
2

  exp
    10.1
 8.314  J/mol K    288.15 K  
1

Therefore,

[ 1  1 f1  10.1 65.0  kPa   657  kPa 

108
8.71
The problem statement provides the following information:

â â
x aá  0.987 x a  0.0013 xbá  0.013 xb  0.9987

At equilibrium:

 xa 
xaá ã aá  xaâ ã aâ  ln   ln    ln  
 x  a a
 a 
 x 
 ln b   ln    ln  
â â
xbá ã bá  xb ã b
 x  b b
 b 

Use the composition data provided in the problem statement and the expressions from Table 7.1:

 0.987 
ln  
1
 0.0013  8.314  298.15 
     
   
 A  3B 0.9987 2  0.013 2  4B 0.9987 3  0.013 3 
   

 0.013 
ln  
1
 0.9987  8.314  298.15 
     
  
 A  3B 0.0013 2  0.987 2  4B 0.0013 3  0.987 3 
   

Solve simultaneously:

 J 
A  13700 
 mol 
 J 
B  2798.5 
 mol 

109
8.72
The instability condition is given by Equation 8.33:

  2g 
 2  0 (8.55)
  xa T ,P

If the nonideality is described by the Wilson equation, we can write the Gibbs energy as:

g  x ag a  xb g b  RT  xa ln xa  x b ln x b   RT  xa ln  xa   ab xb   x b ln x b   ba x a 

Differentiating gives:

 xa 
ln  xa   ab xb  
 g    xa   ab xb  
   ga  gb  RT  ln xa 1 xb ln xb 1  RT
  xa T ,P  xb 
 ln  xb   ba xa   
  xb   ba xa  
A second differentiation gives

  2g  1 1  2 xa 2 xb 
 2  RT     RT     
  xa T ,P  xa x b    xa   ab xb   xa   ab xb   xb   ba xa   xb   ba xa  
2 2

Rearrangement and application of Equation 8.33 gives

  2g   1 1  1 1 
 2  RT  
   
      0
  xa T ,P  xa  xa   ab xb    xb  xb   ba xa  

Since  ab and  ba are greater than zero, the second term in parenthesis is always less than the
first and the above expression can never be satisfied. Hence, the Wilson equation is incapable of
describing the instability of partially miscible liquids.

110
8.73
To solve for the equilibrium mole fractions of a and b in each phase, we set the fugacities of
liquid phase equal to liquid phase . Since the reference state fugacities are equal, we get:

x   
a  a  xa  a (1)
and
x   
b  b  xb  b (2)

We can substitute in the activity coefficient expressions for the three-suffix Margules equation,
 
and use x  
b  1 xa and x b  1 x a to get:

 A  3B   A  3B 

x a exp
 RT
  2 4B
1 x a 
RT
 3 
1 x a  xa exp


 RT
  2 4B
1 x a 
RT
 3
1 xa 

    (3)

and
 A  3B  2 4B  3  A  3B   2 4B  3
 

1 x a exp
 RT
xa 
RT
  
xa  1 x a exp

  
 RT
xa 
RT

xa 

    (4)

If we plot the quantities x a a and 1 x a  b vs. xa, the solution occurs at the two compositions
of xa where Equations 3 and 4 are simultaneously satisfied This solution is shown in the
following figure. If you examine the figure, you will see that the solution illustrated is unique
and no other values of xa will work.
1.4

1.3 x a a
x b b
1.2

xii 1.1

1 x   
b  b  xb  b  0.960

0.9
   
x a  a  xa  a  0.877
0.8

0.7
0 0.2 0.4 0.6 0.8 1
x a  0.051 
x a  0.846 xa

The graphical solution is presented above. It gives:

x a a  xa  a  0.877

x b  b  x b  b  0.960

111
x
a  0.846

x a  0.051

Alternatively, we can plot axavs. bxb to give the same result:

1.4
x a a
1.2
xa = 1 x
a  0.846
1

x   
a  a  xa  a  0.877
0.8
x a  0.051
0.6

0.4

0.2
xa = 0
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4

x   
b  b  xb  b  0.960
x b b

112
8.74
The mixture will split into two phases when

2g
0
xa2

The expression for Gibbs energy is


g  xa ga  1 xa  gb  RT  xa ln xa  1 xa  ln 1 xa   xa 1 xa   A  B  2xa 1
Differentiating twice with respect to xa gives

2g  1 1 
 RT     2 A  6 Bxb  xa  0
xa2  xa xb 

Substituting values from the problem statement and solving provides the following range:

0.155  xa  0.665

113
8.75
(a)

Looking at the chemical species involved:

 Furfural: C5H4O2
 Isobutene: C4H8
They are dissimilar and the activity coefficients are not likely to be symmetric.

Use the three-suffix Margules model.

ln 3 4

3 4

ln 3 4

3 4

The phases are in equilibrium so we can say:

So, we have two equations and two unknowns, plug in values for mole fractions and solve for A
and B:

3 4 3 4
∙ ∙

3 4 3 4
∙ ∙

J
7900
mol

114
J
560
mol

(b)
The like interactions are stronger as shown with a large positive A.

(c)
exp

Calculate the saturation pressure of species a given our system temperature using the Antoine
Equation.

2032.76
exp 8.9179 9.31 bar
337.15 33.15

We can calculate the saturation fugacity coefficient using the generalized correlations.

0.837

The Poynting Correction is negligible at our “low” (below 100 bar) system and saturation
pressure. Solve for the activity coefficient of species a in the beta phase by using the earlier
developed equations.

3 4
exp 8.34
Next:

0.113 8.34 0.837 9.31 7.35 bar

115
8.78
We can use the three-suffix Margules equation to model the non-ideality. At infinite dilution:

RT ln  
a  RT ln 7.02  A  B

RT ln  b  RT ln 72.37  A  B

Solving simultaneously,

 J   J 
A  7720 B  2890
mol  mol 

To solve for the equilibrium mole fractions of a and b in each phase, we set the fugacities
of liquid phase , liquid phase and the vapor phase equal.

x a a Pasat  x a  a Pasat  y aP (1)


and
x b b Pbsat  x b  b Pbsat  y bP (2)

We can substitute in the activity coefficient expressions for the three-suffix Margules equation,
 
Equations 7.128 and 7.129, use x  
b  1 xa and x b  1 x a , and divide by Pi , to get:
sat

 A  3B 
 2 4B
    xa exp ART
 3B 
1 xa    3 y P

  3 2 4B
x a exp 1 x a  1 x a  1 xa  asat
 RT RT RT  Pa
(3)
and

 A  3B  2 4B  3  A  3B   2 4B  3


 

1 x a exp
 RT
xa   
x
RT a    

 1 x

a 
exp
 RT
xa   
x
RT a    PybbsatP
(4)

If we plot the quantities x a  a and 1 x a  b vs. xa, the solution for the composition in the two
liquid phases occurs at the two compositions of xa where Equations 3 and 4 are simultaneously
satisfied. This solution is shown in the following figure. If you examine the figure, you will see
that the solution illustrated is unique and no other values of xa will work.

116
2.3 x b b

xii

1.9

x a a

1.5

1.1
x   
a  a  xa  a  0.987

x   
b  b  xb  b  0.860
0.7
0 0.2 0.4 0.6 0.8 1 xa
 
x a  0.145 x a  0.986

The graphical solution is presented above. It gives:

x a  a  x a  a  0.987

x b b  xb  b  0.860

x
a  0.986

x a  0.145

Alternatively, in analogy to Figure E8.14B, we can plot axavs. bxb to give the same result:

x a a
1.6

x
a  0.986
1.2
xa = 1
x   
a  a  xa  a  0.877
0.8
x a  0.145

0.4

xa = 0
0
0 0.5 1 1.5 2 2.5
x b b
x   
b  b  xb  b  0.860

117
To solve for pressure we can add together Equations 1 and 2 and substitute the appropriate
activity coefficient expression to give:

(5)
P  x a  a Pasat  x b  b Pbsat

At 25 oC, the saturation pressure for water and 1-butanol are 3,169 Pa (from steam tables) and
875 Pa (from problem statement). Solving gives:

P = 3880 Pa

The mole fraction in vapor is given by solving equation 1:

118
8.79

(a)
The two-suffix Margules equation is reasonable since the system is symmetric, i.e.:

x1  x2 and x1  x2

(b)
Equate the fugacities of species 1.

fˆ1  fˆ1
x1  1  x1  1

For the two-suffix Margules equation,

 A  2  A  2
1  exp 
 RT
 x2   1  exp 
 RT
 x2  
Now we can solve for A:

 A  2   A  2
x1 exp 
 RT
 x2    x1 exp 
  RT
 x2  
A  8162 J/mol

(c)
We will assume that the vapor behaves ideally. For phase equilibrium to exist

y1 P  x1  1 P1sat
y 2 P  x2  2 P2sat

Therefore,

P  x1  1 P1sat  x2  2 P2sat


  A 2    A 2 
P   0.9098  exp   0.0902    0.888 bar    0.0902  exp   0.9098  1.223 bar 
  RT    RT 
P  1.97 bar 

(d)
The composition of the vapor phase is found as follows:

119
0.9098 exp  A
0.0902 2  0.888 bar 
  RT 
y1   0.42
1.97 bar
y 2  1  y1  0.58

120
8.80

(a)
Crystallization works better because the difference in freezing points is much greater than the
difference in boiling points. Because the greatest difference between boiling points is 5.5 K,
distillation is an ineffective separation technique.

(b)
(i).
For phase equilibrium in solids, we have the following relationship

 h fus , i s fus , i
lnxi  i   
RT R

Since we are neglecting c P and assuming the solution behaves ideally, the relationship
simplifies to

Tm h fus , i
T
h fus , i  RTm ln xi 

For each isomer of xylene, we calculate the temperature as follows

 248.1 K 13608  J/mol


To   205 K
13608  J/mol  8.314  J/mol K   248.1K  ln  0.25
 225.4 K  11577  J/mol
Tm   202.7 K
11577  J/mol  8.314  J/mol  K   225.4 K  ln  0.5
 286.6 K  17125  J/mol
Tp   240.3 K
17125  J/mol  8.314  J/mol K   286.6 K  ln  0.25

(ii).
First, assume that the ortho- and meta- isomers do not crystallize and calculate the mole fractions
after half of the para-xylene has crystallized.

x1  0.286 x 2  0.571 x3  0.143

Calculate the temperature as before, but use the new mole fractions.

To  208.5 K
Tm  206.7 K
T p  225.5 K

121
Since the temperature for the para-xylene is greater than the other two temperatures, the other
isomers will not crystallize, and the value for Tp is our final answer.

(iii).
Create expressions for the mole fractions of each isomer where only x moles of para-xylene
crystallizes.

1 2 x
xo  xm  xp 
3 x 3 x 3 x

Assume that o-xylene will crystallize before m-xylene. Therefore, we need to find the
temperature at which both species 1 and species 3 crystallize. This is the limit where only para-
xylene crystallizes.

 248.1 K  13608  J/mol



 1 
13608  J/mol  8.314  J/mol  K   248.1K  ln  
 3 x 
 286.6 K  17125  J/mol
 x 
17125  J/mol  8.314  J/mol  K   286.6 K  ln  
 3 x 

Solve for x:

x  0.242

Therefore,

x p  0.075
and
T  210.6 K

To confirm that none of the m-xylene crystallizes, calculate its freezing temperature for this
composition. We obtain

Tm  209.1 K

Since its temperature is lower, the lowest system temperature where only p-xylene crystallizes is
210.6 K. 75.8 % of the p-xylene has crystallized at this temperature.

(c)
(i).
Assume that both the liquid and vapor are ideal. We calculate vapor mole fractions as follows

122
xi Pi sat x P sat
yi   i i sat
P  xi Pi
For the liquid solution of Part (b), we obtain

 0.25 89.7  kPa 


yo   0.22
 0.25 89.7  kPa    0.5 103.5  kPa    0.25 105.5  kPa 
 0.5 103.5  kPa 
ym   0.52
 0.25 89.7  kPa    0.5 103.5  kPa    0.25 105.5  kPa 
 0.25 105.5  kPa 
ym   0.26
 0.25 89.7  kPa    0.5 103.5  kPa    0.25 105.5  kPa 

(ii).
A spreadsheet can be set up for this calculation. One row in the spreadsheet acts as a stage in the
column. The vapor mole fraction of one stage acts as the liquid mole fraction for the next stage
(the next row below). The range of the spreadsheet is extended until the desired purity in the
vapor phase is obtained. The first 10 stages in the spreadsheet are provided below.

Constants
T (ºC) Posat (kPa) Pmsat (kPa) Ppsat (kPa)
140 89.7 103.5 105.5

Liquid Mole Fractions Vapor Mole Fractions


Stage # xo xm xp P (kPa) yo ym yp
1 0.25 0.50 0.25 100.55 0.22 0.51 0.26
2 0.22 0.51 0.26 100.95 0.20 0.53 0.27
3 0.20 0.53 0.27 101.31 0.18 0.54 0.29
4 0.18 0.54 0.29 101.65 0.15 0.55 0.30
5 0.15 0.55 0.30 101.96 0.14 0.56 0.31
6 0.14 0.56 0.31 102.23 0.12 0.56 0.32
7 0.12 0.56 0.32 102.48 0.10 0.57 0.33
8 0.10 0.57 0.33 102.71 0.09 0.57 0.33
9 0.09 0.57 0.33 102.91 0.08 0.58 0.34
10 0.08 0.58 0.34 103.09 0.07 0.58 0.35

To obtain a purity of 90% p-xylene, 149 stages are required.

123
8.81
The problem statement does not specify whether beer is 4% ethanol (1) by mass or volume.
Assume it is based on mass. We need to first convert the mass fraction to a mole fraction.

xˆ1 0.04
MW 1 46.0684
x1    0.016
xˆ1 xˆ 2 0.04 0.96
 
MW 1 MW 2 46.0684 18.0148
Now, we can use the Equation 8.57 to calculate the new freezing point.

x1 RTm2
T  Tm 
h fus ,Tm

Substituting the proper values into Equation 8.57, we obtain

 0.016 8.314  J/mol  K   273.15  K 


2

T  273.15 K 
 6.01103  J/mol
T  271.5 K

Therefore, your freezer can be set to -1.65 ºC (29.0 ºF).

124
8.82
Since the bismuth and cadmium are completely immiscible, we know

f as  xa ã a f al
f bs  xb ã b f bl

sl , we can manipulate Equation 8.45 to obtain


Neglecting c P

h fus ,Tm  1 1 
ln  a       ln  xa 
R  T Tm 
h fus ,Tm  1 1 
ln  b       ln  xb 
R  T Tm 

We can assume the three-suffix Margules equation accurately represents the solution. The
expressions for the activity coefficients are

ln a  
 A  3B xb2  4 Bxb3
RT

ln b  
 A  3B xa2  4 Bxa3
RT

Using the data provided in the problem statement and the eutectic point composition, we obtain
the following equations

 A  3B  0.55  4B  0.55  10460  1  1   ln 0.45


2 3

   
8.314  417.15 8.314  417.15 544.15 
 A  3B  0.45  4B  0.45  6100  1  1   ln 0.55
2 3

   
8.314  417.15 8.314  417.15 594.15 

Solving simultaneously, we get

 J 
A  1157 
 mol 
 J 
B  90.3 
 mol 

125
8.83
Since the solids are completely immiscible, we know

f as  xa a f al
f bs  xb b f bl

sl is negligible, we can manipulate Equation 8.45 to obtain


Assuming c P

h fus ,Tm  1 1 
lnx a  a     
R  T Tm 
h fus ,Tm  1 1 
ln xb  b     
R  T Tm 

Now determine the activity coefficients from the excess Gibbs energy

RT ln  a  GaE  60001  0.0005T xb2


RT ln  b  GbE  60001  0.0005T 1  xb 2

Substitute these expressions into the equations above:

 60001  0.0005T x 2  h fus,T , a  1 1 


ln1  xb    b
 m
  
 RT  R 
 T Tm, a 
 
 60001  0.0005T 1  x 2  h fus ,T , b  1 1 
ln  xb    b
 m
  
 RT  R  T Tm, b 

Above, we have two equations for two unknowns: T and xb. Solving simultaneously, we get

T  656 K
xb  0.66

Therefore, the composition at the eutectic point is

xa  0.34
xb  0.66

126
8.84
Let the subscript “1” designate antimony and “2” designate lead. At the eutectic temperature, a
liquid solution, an antimony-rich solid solution, and lead-rich solid solution are in equilibrium.
Assume the solid solutions at the eutectic temperature exhibit ideal solution behavior.
Furthermore, we can assume the liquid solution is ideal since it is 11.2% antimony. Therefore,

1  1 1  1
&
2  1 2 1

We can also assume the change in heat capacity between solid and liquid phases is negligible.
Thus, the relationship required to find the composition of the solid solution reduces to

 x  h fus, 2  1 1 
ln 2   
 X2  R  T Tm,2 

where T is the eutectic temperature and X2 is the mole fraction of lead in the lead-rich phase.
Find the mole fractions at the eutectic point from the mass fractions.

x1  0.177
x 2  0.823

Now, solve the two expressions for the solid compositions:

0.823
X2   0.955
  5100  1 1 
exp    
 8.314  524.15 600.65 
X 1  0.045

127
8.85

(a)
The eutectic point defines the lowest temperature where a binary mixture can exist entirely as a
liquid. At the eutectic point

T  780 º C
zsilver  0.6
zcopper  0.4

(b)
The lowest mole percentage of silver in the solid silver phase is approximately 0.86. Therefore,

X copper  0.14

This occurs at a temperature of 780 ºC.

(c)
For this composition and temperature, two phases form. They are:

 a liquid solution
xsilver  0.55
xcopper  0.45
 a solid solution
X silver  0.05
X copper  0.95

To calculate the number of moles in each phase, we can use mole balances or the lever rule.
Mole balances are shown below, where nliq is the number of moles of the liquid phase and nsol is
the number of moles in the solid phase.

nliq  nsol  5 mol


 0.55 nliq   0.05 nsol 1 mol
Solve simultaneously:

nliq  1.5 mol


nsol  3.5 mol

128
8.86

129
8.87
We have a system with water (1) and ethanol (2). When we add the ethanol, the entropy of
the liquid increases, lowering its Gibbs energy relative to the Gibbs energy of the pure solid:

1l  g1s

Thus, the solid ice will melt. However, it takes energy for the solid to melt. While a fraction
of that comes from the exothermic enthalpy of mixing, the remainder comes from the
molecular kinetic energy of the molecules in the system. Thus the temperature will lower. If
it all melts, the weight fraction of ethanol will be 0.33 and we can use the phase diagram in
the problem to see we only have one liquid phase to -20 oC

To see if we end up above -20 oC, we perform an energy balance to get

H  0  H sensible  H latent  H mix

with

H sensible  n1l cPl ,1 (T f  0)  n2l cPl ,2 (T f  0)  n1s cPs ,1 (T f  0)

and
H latent   n1s ,0  n1s  h1fus

where n1s ,0 is the initial number of moles of solid ice (55.6 kg) Finally, we have
H mix  nl x1 x2 R  190.0  214.7  x2  x1   419.4  x2  x1   383.3  x2  x1   235.4  x2  x1  
2 3 4
 

where we convert from mass fraction of the liquid to the mole fraction of liquid using:

w1
 MW 1
x1 
w1 w2

 MW 1  MW 2
We now can use a guess and check procedure combining the energy balance with the phase
diagram in the problem statement. For example, if we assume ALL the ice melts, we apply
the energy balance to get:

Tf = -26.5 oC

An excerpt from the spreadsheet to perform this calculation is shown below

130
H2O ETOH fraction melt
CP/R 9.069 13.59 1
M 2 1 kg
n 111.11 21.74 mol
MW 0.018 0.046
x 0.56 0.44
w 0.33 0.67

Dhmelt 3.34E+05 J
deltahmix ‐349.48 J/mol
Dhmix ‐4.64E+04 J
Delta H 2.87E+05 J
Tf ‐26.5 C

Since that is in the 2 phase region, not all the ice melts. We need to iterate to where the
fraction of ice that melts matches the amount of solid that remains using the lever rule in the
phase diagram below. This occurs where

mice = 0.15 kg and w1 = 0.35

Tf = -22 oC

The final state is shown on the phase diagram below. It is mostly, but not all, liquid.

131
8.88
We can get a first approximation of the heat of fusion and melting temperature by assuming that
sl is negligible. With these assumptions, Equation 8.47 reduces to
the solutions are ideal and c P

x  h fus ,Tm  1 1 
ln i     
 Xi  R  T Tm 

 x  h fus,Tm h fus ,Tm


Now, plot ln i  vs. 1/T. The slope is , and the y-intercept is  .
 Xi  R RTm

ln(x Fe /X Fe ) vs. 1/T

0
ln(xFe/XFe )

-0.05
y = -807x + 0.4272
R2 = 0.9671
-0.1

-0.15
0.0005 0.00055 0.0006 0.00065 0.0007 0.00075
-1
1/T [K ]

From the line of best fit on the above graph,

h fus ,Tm  J 
 807 K   h fus ,Tm  6709.4  
R  mol 

h fus ,Tm
  0.4272  Tm  1889 K 
RTm

To obtain more accurate values of the heat of fusion and melting temperature, do not assume the
sl . Equation 8.47 becomes
solutions are ideal. We can still neglect c P

 x  h fus ,Tm 1 1 
ln i i     
 X i i  R  T Tm 

132
Now, copy the given data table into a spreadsheet and create the following headings in the
spreadsheet.

T [oC] T[K] XC xC ln(xFe/XFe) Fe Fe ln(Fe/Fe) RHS (LHS-RHS)2

The first five columns are self-explanatory. The cells in the remaining columns contain the
following formulas:

 Ax 2 
Fe:  exp c 
 RT 
 
Note: T references the cell in the same row containing the experimental
temperature (in Kelvin). xc references the corresponding experimental value in
the xc column. R is 8.314. A is a reference to a cell containing an initial guess for
A. Use 0 for an initial guess.

A X2
Fe:  exp sol c 
 RT 
 
Note: T references the cell in the same row containing the experimental
temperature (in Kelvin). Xc references the corresponding experimental value in
the Xc column. R is 8.314. Asol is a reference to a cell containing an initial guess
for Asol. Use 0 for an initial guess.

ln(Fe/Fe):  ln  Fe / Fe 

h fus , T m  1 1 
RHS:    
R  T T m 
Note: T references the cell in the same row containing the experimental
temperature (in Kelvin). R is 8.314. h fus ,Tm is a reference to a cell containing
the first approximation obtained from the above graph. Tm is a reference to a cell
containing the first approximation obtained earlier.

(LHS-RHS)2:  ln  xFe / X Fe   ln  Fe /  Fe   RHS 


2

Create a spreadsheet in which all of the data are used in the above formulas. In once cell,
calculate the sum of all the entries in the (LHS-RHS)2 column. Go to the spreadsheet’s “Solver”
function and choose to minimize the sum of the (LHS-RHS)2 column by changing the values of
the cells that contain A, Asol, Tm, and h fus ,Tm . The following spreadsheet results from
following this procedure:

133
(LHS-
T [K] XC xC ln(xFe/XFe) Fe Fe ln(Fe/Fe) RHS
RHS)^2
1421.15 0.1 0.2092 -0.129 0.845 1.020 -0.189 -0.323 25.101
1427.15 0.09 0.2072 -0.138 0.848 1.016 -0.181 -0.317 5.185
1473.15 0.0877 0.1906 -0.120 0.874 1.015 -0.150 -0.268 2.031
1523.15 0.0718 0.1689 -0.110 0.902 1.010 -0.112 -0.219 14.186
1573.15 0.0613 0.145 -0.093 0.929 1.007 -0.080 -0.173 0.302
1623.15 0.0475 0.1179 -0.077 0.954 1.004 -0.051 -0.130 3.751
1673.15 0.0333 0.0891 -0.059 0.974 1.002 -0.028 -0.089 2.480
1723.15 0.0196 0.057 -0.039 0.990 1.001 -0.011 -0.051 0.530
1768.15 0.0079 0.0248 -0.017 0.998 1.000 -0.002 -0.018 1.543
SUM 55.109

deltahfus: -18376.7 J/mol


Tm : 1793.9 K
A: -45578.7 J/mol
Asol: 23622.1 J/mol

Therefore,

h fus,Tm  18.4  kJ/mol


Tm 1794  K 

We must be careful using the solver function in Excel for this non-linear set of equations. In
fact, the solution above represents a local minimum. Other numerical techniques can be
employed to solve this problem. For example, one technique provides a sum of the error
function as low as 19.62. Here are the resulting values for that technique:

deltahfus -15,501 J/mol


Tm 1799.3 K
A -19174 J/mol
Asol 88446 J/mol

Can you obtain values close to these on your own?

134
8.89
We have the following relationship:

hvap  1 
 ln 1  xurea  
1
  
R  T Tboil 

where

T  329.2 K
Tboil  329.44 K
0.009  kg urea 
xurea   0.0089
1  kg acetone  0.009  kg urea 

We can rearrange the equation to obtain

 R ln 1  xurea 
hvap 
1 1 
  
 T Tboil 

which upon substitution yields

 kJ 
hvap  33.6  
 mol 

135
8.90
We can use Equation 8.56 to solve for the required mole fraction:

h fus ,Tm  1 1 
lnxa   ln 1  xb     
R  T Tm 
(Note: xb is the mole fraction of ethylene glycol.)

Substitute numerical values and solve for xb :

  J  
  6010  mol  
 1  exp     1

1 
xb  
  J   263.15 K 273.15 K 
 8.314  mol  K  
 
xb  0.096

Now convert the mole fraction to a mass fraction:

xb  MW  b
wb 
xa  MW  a  xb  MW  b
0.096  62.0674 
wb   0.268
0.904 18.0148  0.096  62.0674

Since the density of ethylene glycol is equal to the density of water, the required mass fraction is
equal to the fraction by volume. To prevent freezing at -10 ºC, the water in your radiator must
contain 26.8% antifreeze by volume.

136
8.91
In a reverse osmosis apparatus, if a pressure greater than P    is applied to the compartment
containing sea water, the pure water will be forced into the compartment with pure water, while
the salt is left behind. However, the problem statement doesn’t provide an atmospheric pressure,
so we will assume it is 1 bar. A temperature of 25 ºC can also be used. To calculate the osmotic
pressure,  , use Equation 8.60:

x RT
 b
va
(Note: Species “a” is water and species “b” is the salt.)

The required information and method to calculate xb is provided in Example 8.23.

xb  0.022

Since the mole fraction of salt in seawater is relatively low, we can set the molar volume of
seawater at 25 ºC equal to the molar volume of pure water at 25 ºC. Assuming that the molar
volume of water does not change much with respect to pressure, estimate the molar volume of
pure water from saturation steam tables:

 m3 
vˆasat  0.001003   (Saturated water at 25 ºC)
 kg 
  m 
3  m3 
va   0.001003    0.0180148  kg/mol  1.807 10 
5

  kg   mol 

Now substitute numerical values and evaluate:

0.022  8.314  J 


298.15 K 
  mol  K  
  3.02  10 6 Pa  30.2 bar 
 m3 
1.807  10  5  
 mol 

Therefore, the minimum pressure required for reverse osmosis of seawater is

P    31.2 bar 

137
8.92
First, calculate the mole fractions of water (species a) and sucrose (species b).

500  g
na   27.8  mol H 2 O
18.0148 g/mol
0.5 g
nb   0.0015  mol sucrose
342.3 g/mol

27.8 mol
 xa   0.9999
27.8 mol  0.0015 mol
xb  1  xa  0.0001

Since the solution is dilute, we can use Equation 8.60 to calculate the osmotic pressure.

x RT
 b
va

We have all of the necessary values to calculate the osmotic pressure except the molar volume of
water at 25 ºC and 1 bar. Since liquid molar volumes do not change much with pressure, we can
use the molar volume of saturated water at 25 ºC. From the steam tables

 m3 
v̂asat  0.001003  
 kg 

Therefore,

  m 3   m3 
   0.0180148  kg/mol  1.807 10 
5
va   0.001003 
  kg   mol 

Now substitute values into Equation 8.60 and evaluate:

  J 
 0.0001 8.314    298.15  K 
 mol  K 

 m3 
1.807 10 5  
 mol 
  1.37 10 4  Pa 

138
8.93

Problem 8.55
Isothermal data are given; therefore, select the Isothermal option when finding the activity
coefficient parameters with ThermoSolver. Enter the data into a new data table. We can then
calculate parameters for any activity coefficient model using the objective function of our choice.
The parameters provided below were obtained using the pressure objective function (OFP).

2 Suffix Margules Three Suffix Margules van Laar Wilson NRTL


A A B A B 12 21 G12 G21 12 21
-2329.29 -2344.44 353.62 -2740.21 -2032.44 1.125 1.964 1.104 0.858 2.418 -3.735

Problem 8.56
Isobaric data are given; therefore, select the Isobaric option when finding the activity coefficient
parameters with ThermoSolver. Enter the data into a new data table. The parameters provided
below were obtained using the pressure objective function (OFP).

2 Suffix Margules Three Suffix Margules van Laar Wilson NRTL


A/RT A/RT B/RT A/RT B/RT 12 21 G12 G21 12 21
1.949 1.839 -0.383 2.254 1.547 0.18 0.354 1.68 1.211 1.108 0.408

139
8.94
Since we know the vapor composition and we want the liquid composition and dew point
temperature at various pressures, this problem corresponds to quadrant IV shown in Figure 8.1.

(a)
To solve using Raoult’s law, make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine 


Fugacity Coefficient:  1
Activity Coefficients:  1

We obtain

P T x1 x2 x3 x4
(bar) (ºC) n-Hexane Cyclohexane Benzene Toluene
1 91.0 0.10 0.18 0.18 0.54
20 237.5 0.14 0.22 0.20 0.44

(b)
Make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine   sat  exp...


Fugacity Coefficient:  1
Activity Coefficients: Multicomponent Wilson

We obtain

P T x1 x2 x3 x4
(bar) (ºC) n-Hexane Cyclohexane Benzene Toluene
1 91.0 0.10 0.18 0.18 0.54
20 263.5 0.19 0.22 0.22 0.38

(c)
Make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine   sat  exp...


Fugacity Coefficient: ̂ Peng Robinson 
Activity Coefficients: Multicomponent Wilson

We obtain

P T x1 x2 x3 x4
(bar) (ºC) n-Hexane Cyclohexane Benzene Toluene
1 88.8 0.09 0.17 0.19 0.55
20 230.5 0.16 0.22 0.23 0.39

140
To compare the results from Parts a-c, the ranges of the compositions and dew-point
temperatures can be computed.

Ranges
P (bar) T (ºC) x1 x2 x3 x4
1 2.2 0.01 0.01 0.01 0.01
20 33.0 0.05 0.00 0.03 0.06

Clearly, the range is greater for all of the quantities when the pressure is 20 bar. This is expected
since greater non-ideality exists at higher pressures.

141
8.95
Since we know the liquid composition and want the bubble-point temperature and vapor
composition at various pressures, this problem corresponds to quadrant III in Figure 8.1.

(a)
To solve using Raoult’s law, make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine 


Fugacity Coefficient:  1
Activity Coefficients:  1

We obtain

P T y1 y2 y3 y4
(bar) (ºC) n-Hexane Cyclohexane Benzene Toluene
1 83.4 0.31 0.27 0.28 0.13
20 231.4 0.27 0.26 0.28 0.19

(b)
Make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine   sat  exp...


Fugacity Coefficient:  1
Activity Coefficients: Multicomponent Wilson

We obtain

P T y1 y2 y3 y4
(bar) (ºC) n-Hexane Cyclohexane Benzene Toluene
1 81.5 0.32 0.28 0.28 0.13
20 260.6 0.20 0.28 0.28 0.24

(c)
Make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine   sat  exp...


Fugacity Coefficient: ̂ Peng Robinson 
Activity Coefficients: Multicomponent Wilson

We obtain

P T y1 y2 y3 y4
(bar) (ºC) n-Hexane Cyclohexane Benzene Toluene
1 80.4 0.32 0.28 0.28 0.13
20 226.1 0.24 0.27 0.27 0.22

142
To compare the results from Parts a-c, the range of the three values for the bubble-point
temperature and each composition can be computed.

Range
P (bar) T (ºC) y1 y2 y3 y4
1 3.0 0.01 0.01 0.00 0.00
20 34.5 0.07 0.02 0.01 0.05

The standard deviation is greater for all of the quantities when the pressure is 20 bar. This is
expected since greater non-ideality exists at higher pressures.

143
8.96
Since we know the vapor composition and we want the liquid composition and dew-point
pressure at various temperatures, this problem corresponds to quadrant I shown in Figure 8.1.

(a)
To solve using Raoult’s law, make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine 


Fugacity Coefficient:  1
Activity Coefficients:  1

We obtain

T P x1 x2 x3
(ºC) (bar) Methanol Acetone n-Hexane
40 0.416 0.29 0.26 0.45
200 23.86 0.15 0.31 0.55

(b)
Make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine   sat  exp...


Fugacity Coefficient:  1
Activity Coefficients: Multicomponent Wilson

We obtain

T P x1 x2 x3
(ºC) (bar) Methanol Acetone n-Hexane
40 0.745 0.17 0.29 0.54
200 25.77 0.06 0.27 0.68

(c)
Make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine   sat  exp...


Fugacity Coefficient: ̂ Peng Robinson 
Activity Coefficients: Multicomponent Wilson

We obtain

T P x1 x2 x3
(ºC) (bar) Methanol Acetone n-Hexane
40 0.768 0.19 0.30 0.51
200 230.5 0.16 0.22 0.23

144
The results above are reasonable for 40 oC, but not for 200 oC since the mole fractions do not
sum to one. (It turns out that these equations do not give a solution for the higher temperature).
Thus, for 200 oC, the best answer is given by part b. To compare the results from the parts we
believe, the range of the compositions and dew-point temperatures can be compared.

Ranges
T P x1 x2 x3
40 0.352 0.12 0.04 0.09
200 1.89 0.09 0.04 0.13

145
8.97
Since we know the liquid composition and we want the vapor composition and bubble-point
pressure at various temperatures, this problem corresponds to quadrant II shown in Figure 8.1.

(a)
To solve using Raoult’s law, make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine 


Fugacity Coefficient:  1
Activity Coefficients:  1

We obtain

T P y1 y2 y3
(ºC) (bar) Methanol Acetone n-Hexane
40 0.434 0.20 0.45 0.34
200 26.68 0.38 0.36 0.26

(b)
Make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine   sat  exp...


Fugacity Coefficient:  1
Activity Coefficients: Multicomponent Wilson

We obtain

T P y1 y2 y3
(ºC) (bar) Methanol Acetone n-Hexane
40 0.75 0.26 0.36 0.38
200 32.54 0.45 0.27 0.28

(c)
Make the following selections in ThermoSolver:

Fugacity of Pure Liquid: P sat Antoine   sat  exp...


Fugacity Coefficient: ̂ Peng Robinson 
Activity Coefficients: Multicomponent Wilson

We obtain

T P y1 y2 y3
(ºC) (bar) Methanol Acetone n-Hexane
40 0.771 0.26 0.36 0.39
200 230.5 0.16 0.22 0.23

146
The results above are reasonable for 40 oC, but not for 200 oC since the mole fractions do not
sum to one. (It turns out that these equations do not give a solution for the higher temperature).
Thus, for 200 oC, the best answer is given by part b. To compare the results from the parts we
believe, the range of the compositions and dew-point pressures can be compared.

Ranges
T P y1 y2 y3
40 0.337 0.06 0.09 0.05
200 6.86 0.07 0.09 0.02

147

You might also like